FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 21. Regional Context Coordinating Lead Authors Bruce Hewitson (South Africa), Anthony C. Janetos (USA) Lead Authors Timothy R. Carter (Finland), Filippo Giorgi (Italy), Richard G. Jones (UK), Won-Tae Kwon (Republic of Korea), Linda O. Mearns (USA), E. Lisa F. Schipper (Sweden), Maarten K. van Aalst (Netherlands) Contributing Authors Eren Bilir (USA), Monalisa Chatterjee (USA / India), Grace Redmond (UK), Carol McSweeney (UK), Katharine Mach (USA), Vanessa Schweizer (USA), Luke Wirth (USA), Claire van Wyk (South Africa) Review Editors Thomas Downing (USA), Phil Duffy (USA) Volunteer Chapter Scientist Kristin Kuntz-Duriseti (USA) Contents Executive Summary 21.1. Introduction 21.2. Defining Regional Context 21.2.1. Decision-Making Context 21.2.2. Defining Regions 21.2.3. Introduction to Methods and Information 21.3. Synthesis of Key Regional Issues 21.3.1. Vulnerabilities and Impacts 21.3.1.1. Observed Impacts 21.3.1.2. Future Impacts and Vulnerability 21.3.2. Adaptation 21.3.2.1. Similarities and Differences in Regions 21.3.2.2. Adaptation Examples in Multiple Regions 21.3.2.3. Adaptation Examples in Single Regions 21.3.3. Climate System 21.3.3.1. Global context 21.3.3.2. Dynamically and Statistically Downscaled Climate Projections 21.3.3.3. Projected Changes in Hydroclimatic Regimes, Major Modes of Variability, and Regional Circulations 21.3.3.4. Projected Changes in Extreme Climate Events 21.3.3.5. Projected Changes in Sea Level 21.3.3.6. Projected Changes in Air Quality 21.4. Cross-Regional Phenomena 21.4.1. Trade and Financial Flows 21.4.1.1. International Trade and Emissions 21.4.1.2. Trade and Financial Flows as Factors Influencing Vulnerability 21.4.1.3. Sensitivity of International Trade to Climate 21.4.2. Human Migration Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 21.4.3. Migration of Natural Ecosystems 21.5. Analysis and Reliability of Approaches to Regional Impacts, Adaptation, and Vulnerability Studies 21.5.1. Analyses of Vulnerability and Adaptive Capacity 21.5.1.1. Indicators and Indices 21.5.1.2. Hotspots 21.5.2. Impacts Analyses 21.5.3. Development and Application of Baseline and Scenario Information 21.5.3.1. Baseline Information: Context, Current Status, and Recent Advances 21.5.3.2. Development of Projections and Scenarios 21.5.3.3. Credibility of Projections and Scenarios 21.6. Knowledge Gaps and Research Needs References Chapter Boxes 21-1. A New Framework of Global Scenarios for Regional Assessment 21-2. Summary Regional Climate Projection Information 21-3. Developing Regional Climate Information Relevant to Political and Economic Regions 21-4. Synthesis of Projected Changes in Extremes Related to Temperature and Precipitation Frequently Asked Questions 21.1: How does this report stand alongside previous assessments for informing regional adaptation? 21.2: Do local and regional impacts of climate change affect other parts of the world? 21.3: What regional information should I take into account for climate risk management for the 20 year time horizon? 21.4: Is the highest resolution climate projection the best to use for performing impacts assessments? Executive Summary There has been an evolution in the treatment of regional aspects of climate change in IPCC reports from a patchwork of case examples in early assessments towards recent attempts at a more systematic coverage of regional issues at continental and sub-continental scales (21.2.2). Key topics requiring a regional treatment include: changes in the climate itself and in other aspects of the climate system (such as the cryosphere, oceans, sea level and atmospheric composition), climate change impacts on natural resource sectors and on human activities and infrastructure, factors determining adaptive capacity for adjusting to these impacts, emissions of greenhouse gases and aerosols and their cycling through the Earth system, and human responses to climate change through mitigation and adaptation. A good understanding of decision-making contexts is essential to define the type and scale of information on climate change related risks required from physical climate science and impacts, adaptation and vulnerability assessments (21.2.1) (high confidence). This is a general issue for all impacts, adaptation and vulnerability assessments, but is especially important in the context of regional issues. Many studies still rely on global datasets, models and assessment methods to inform regional decisions. However, tailored regional approaches are often more effective in accounting for variations in trans-national, national and local decision-making contexts, as well as across different groups of stakeholders and sectors. There is a growing body of literature offering guidance on how to provide the most relevant climate risk information to suit specific decision-making scales and processes. A greater range of regional scale climate information is now available which provides a more coherent picture of past and future regional changes with associated uncertainties (21.3.3). More targeted analyses of reference and projected climate information for impact assessment studies have been carried out. Leading messages include: Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Significant improvements have been made in the amount and quality of climate data that are available for establishing baseline reference states of climate-sensitive systems (21.5.3.1). These include new and improved observational datasets, rescue and digitisation of historical datasets, and a range of improved global reconstructions of weather sequences. A larger set of global and regional (both dynamical and statistical) model projections allow a better characterization of ranges of plausible climate futures than in the AR4 (21.3.3), and more methods are available to produce regional probabilistic projections of changes for use in IAV assessment work (21.5.3). Better process understanding would strengthen the emerging messages on future climate change where there remains significant regional variation in their reliability (21.3.3). Confidence in past climate trends has different regional variability and in many regions there is higher confidence in future changes, often due to a lack of evidence on observed changes (21.3, Box 21-4). In spite of improvements, the available information is limited by the lack of comprehensive observations of regional climate, or analyses of these, and different levels of confidence in projected climate change (high confidence). Some trends that are of particular significance for regional impacts and adaptation include (21.3.3.1; WG I SPM): The globally averaged combined land and ocean surface temperature data show a warming of 0.85 [0.65 to 1.06] °C, over the period 1880 2012. There is regional variation in the global trend, but overall the entire globe has warmed during the period 1901-2012. (WGI SPM) Future warming is very likely to be larger over land areas than over oceans. (WGI SPM) Averaged over mid-latitude land areas, precipitation has increased since 1901(medium confidence before and high confidence after 1951), but for other regions there is low confidence in the assessment of precipitation trends (WGI SPM). There are likely more land regions where the number of heavy precipitation events has increased than where it has decreased. The frequency or intensity of heavy precipitation events has likely increased in North America and Europe. In other continents, confidence in changes in heavy precipitation events is at most medium. The frequency and intensity of drought has likely increased in the Mediterranean and West Africa and likely decreased in central North America and north-west Australia. The annual mean Arctic sea ice extent decreased over the period 1979 2012 with a rate that was very likely in the range 3.5 to 4.1% per decade. Climate models indicate a nearly ice-free Arctic Ocean in September before mid-century is likely under the high forcing scenario RCP8.5 (medium confidence). The average rate of ice loss from glaciers worldwide, excluding those near the Greenland and Antarctic ice sheets, was very likely 275 [140 to 410] Gt yr-1 over the period 1993-2009. By the end of the 21st century, the volume of glaciers (excluding those near the Antarctic ice sheet), is projected to decrease by 15 to 55% for RCP2.6, and by 35 to 85% for RCP8.5, relative to 1986-2005 (medium confidence). The rate of global mean sea-level rise during the 21st century is very likely to exceed the rate observed during 1971 2010, under all RCP scenarios (21.3.3.5; WG I SPM). By the end of the 21st century it is very likely that sea level will rise in more than about 95% of the ocean area, with about 70% of the global coastlines projected to experience a sea level change within 20% of the global mean change. Sea-level rise along coasts will also be a function of local and regional conditions, including land subsidence or uplift and patterns of development near the coast. There is substantial regional variation in observations and projections of climate change impacts, both because the impacts themselves vary, and because of unequal research attention (21.3.1). Evidence linking observed impacts on biological, physical and (increasingly) human systems to recent and ongoing regional temperature and (in some cases) precipitation changes have become more compelling since the AR4. This is due both to the greater availability of statistically robust, calibrated satellite records, and to improved reporting from monitoring sites in hitherto under-represented regions, though the disparity still remains large between data rich and data-poor regions. Regional variations in physical impacts such as vegetation changes, sea-level rise, and ocean acidification are increasingly well documented, though their consequences for ecosystems and humans are less well studied or understood. Projections of future impacts rely primarily on a diverse suite of biophysical, economic and integrated models operating from global- to site-scales, though some physical experiments are also conducted to study processes in altered environments. New research initiatives are beginning to exploit the diversity of impact model projections, through cross-scale model inter-comparison exercises. Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 There are large variations in the degree to which adaptation processes, practices and policy have been studied and implemented in different regions (21.3.2) (high confidence). Europe and Australia have had extensive research programs on climate change adaptation, while research in Africa and Asia has been dominated by international partners and relies heavily on case studies of community-based adaptation. National adaptation strategies are common in Europe, and adaptation plans are in place in some cities in Europe, the Americas and Australasia, with agriculture, water and land use management the primary sectors of activity. However, it is still the case that implementation lags behind planning in most regions of the world. Contested definitions and alternative approaches to describing regional vulnerability to climate change pose problems for interpreting vulnerability indicators (21.3.1.2; 21.5.1.1). There are numerous studies that use indicators to define aspects of vulnerability, quantifying these across regional units (e.g. by country or municipality), often weighting and merging them into vulnerability indices and presenting them regionally as maps. However, methods of constructing indices are subjective, often lack transparency and can be difficult to interpret. Moreover, indices commonly combine indicators reflecting current conditions (e.g. of socio-economic capacity) with other indicators describing projected changes (e.g. of future climate or population), and have failed to reflect the dynamic nature of the different indicator variables. Hotspots draw attention, from various perspectives and often controversially, to locations judged to be especially vulnerable to climate change (21.5.1.2). Identifying hotspots is an approach that has been used to indicate locations that stand out in terms of impacts, vulnerability or adaptive capacity (or combinations of these). The approach exists in many fields and the meaning and use of the term hotspots differs, though their purpose is generally to set priorities for policy action or for further research. Hotspots can be very effective as communication tools, but may also suffer from methodological weaknesses. They are often subjectively defined, relationships between indicator variables may be poorly understood and they can be highly scale-dependent. In part due to these ambiguities, there has been controversy surrounding the growing use of hotspots in decision making, particularly in relation to prioritising regions for climate change funding. Cross-regional phenomena can be crucial for understanding the ramifications of climate change at regional scales, and its impacts and policies of response (21.4) (high confidence). These include global trade and international financial transactions, which are linked to climate change as a direct or indirect cause of anthropogenic emissions, as a predisposing factor for regional vulnerability, through their sensitivity to climate trends and extreme climate events, and as an instrument for implementing mitigation and adaptation policies. Migration is also a cross- regional phenomenon, whether of people or of ecosystems, both requiring trans-boundary consideration of their causes, implications and possible interventions to alleviate human suffering and promote biodiversity. Downscaling of global climate reconstructions and models has advanced to bring the climate data to a closer match for the temporal and spatial resolution requirements for assessing many regional impacts, and the application of downscaled climate data has expanded substantially since AR4 (21.3.3; 21.5.3). This information remains weakly coordinated, and current results indicate that high resolution downscaled reconstructions of the current climate can have significant errors. The increase in downscaled data sets has not narrowed the uncertainty range. Integrating these data with historical change and process based understanding remains an important challenge. Characterization of uncertainty in climate change research on regional scales has advanced well beyond quantifying uncertainties in regional climate projections alone, to incorporating uncertainties in simulations of future impacts as well as considering uncertainties in projections of societal vulnerability (21.3.3, 21.5.3, 21.5.1. 21.5.2). In particular, intercomparison studies are now examining the uncertainties in impacts models (e.g., AgMIP and ISIMIP) and combining them with uncertainties in regional climate projections. Some results indicate that a larger portion of the uncertainty in estimates of future impacts can be attributed to the impact models applied rather than to the climate projections assumed. In addition, the deeper uncertainties associated with aspects of defining societal vulnerability to climate change related to the alternative approaches to defining vulnerability are becoming appreciated. As yet there has been little research actively to quantify these uncertainties or to combine them with physical impact and climate uncertainties. Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Studies of multiple stressors and assessments of potential global and regional futures using scenarios with multiple, non-climate elements are becoming increasingly common (21.5.3.2; 21.5.3.3). Non-climatic factors relevant to assessing a system s vulnerability generally involve a complex mix of influences such as environmental changes (e.g. in air, water and soil quality, sea level, resource depletion), land use and land cover changes, and socio-economic changes (e.g. in population, income, technology, education, equity, governance). All of these non- climate factors have important regional variations. There is significant variation in vulnerability due to variability in these factors. 21.1. Introduction This chapter serves as an introduction to Part B of this volume. It provides context for an assessment of regional aspects of climate change in different parts of the world, which are presented in the following nine chapters. While the main focus of those chapters is on the regional dimensions of impacts, adaptation and vulnerability, this chapter also offers links to regional aspects of the physical climate reported by Working Group I and of mitigation analysis reported by Working Group III. The chapter frames the discussion of both global and regional issues in a decision- making context. This context identifies different scales of decisions that are made (e.g. global, international, regional, national, subnational, local) and the different economic or impact sectors that are often the objects of decision-making (e.g. agriculture, water resources, energy). Within this framing, the chapter then provides three levels of synthesis. First there is an evaluation of the state of knowledge of changes in the physical climate system, and associated impacts and vulnerabilities, and the degree of confidence that we have in understanding those on a regional basis as relevant to decision-making. Second, the regional context of the sectoral findings presented in Part A of this volume is discussed. Third, there is an analysis of the regional variation revealed in subsequent chapters of Part B. In so doing, the goal is to examine how the chapters reflect differences or similarities in how decision-making is being addressed by policy and informed by research in different regions of the world, and whether there is commonality of experience among regions that could be useful for enhancing decisions in the future. Having analyzed similarities and differences among IPCC regions, the chapter then discusses trans-regional and cross-regional issues that affect both human systems (e.g. trade and financial flows) and natural systems (e.g. ecosystem migration). Finally, the chapter evaluates methods of assessing regional vulnerabilities and adaptation, impact analyses, and the development and application of baselines and scenarios of the future. These evaluations provide guidance for understanding how such methods might ultimately be enhanced, so that the confidence in research about possible future conditions and consequences might ultimately improve. 21.2. Defining Regional Context The climate system may be global in extent, but its manifestations through atmospheric processes, ocean circulation, bioclimatic zones, daily weather and longer-term climate trends are regional or local in their occurrence, character and implications. Moreover, the decisions that are or could be taken on the basis of climate change science play out on a range of scales, and the relevance and limitations of information on both biophysical impacts and social vulnerability differ strongly from global- to local-scale, and from one region to another. Explicit recognition of geographical diversity is therefore important for any scientific assessment of anthropogenic climate change. The following sections emphasize some of the crucial regional issues to be pursued in Part B of this report. 21.2.1. Decision-Making Context A good understanding of decision-making contexts is essential to define the type and resolution and characteristics of information on climate change related risks required from physical climate science and impacts, adaptation and vulnerability assessments (e.g. IPCC, 2012a). This is a general issue for all impacts, adaptation and vulnerability Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 assessments (cf. the chapters in Part A), but is especially important in the context of regional issues. Many studies still rely on global datasets, models and assessment methods to inform regional decisions. However, tailored regional approaches are often more effective in accounting for variations in trans-national, national and local decision-making contexts, as well as across different groups of stakeholders and sectors. There is a growing body of literature offering guidance on how to provide the most relevant climate risk information to suit specific decision- making scales and processes (e.g., Willows and Connell, 2003; ADB, 2005; Kandlikar et al., 2011). Table 21-1 illustrates the range of actors involved in decision-making to be informed by climate information at different scales in different sectors, ranging from international policy makers and agencies, to national and local government departments, to civil society organizations and the private sector at all levels, all the way to communities and individual households. The table illustrates how policy makers face a dual challenge in achieving policy integration vertically, through multiple levels of governance, and horizontally, across different sectors (policy coherence). [INSERT TABLE 21-1 HERE Table 21-1: Dimensions of the institutions and actors involved in climate change decision-making, including example entries referred to in chapters of this volume. Vertical integration can occur within as well as between levels. Decison-making domains are illustrative. Modified and extended from Mickwitz (2009).] Many climate change risk assessments have traditionally been undertaken either in the context of international climate policy-making (especially the United National Framework Convention on Climate Change UNFCCC), or by (or for) national governments (e.g. CCRA, 2012; SEI, 2009; GCAP, 2011; Roshydromet, 2008). In those cases, climate risk information commonly assumes a central role in the decision-making, for instance to inform mitigation policy, or for plans or projects designed specifically to adapt to a changing climate. In recent years, increasing attention has been paid to more sector- or project-specific risk assessments, intended to guide planning and practice by a range of actors (e.g. Liu et al., 2008; Rosenzweig et al., 2011). In those contexts, climate may often be considered as only one contributor among a much wider set of considerations for a particular decision. In such cases, there is uncertainty about not only the future climate, but also many other aspects of the system at risk. Moreover, while analysts will seek the best available climate risk information to inform the relative costs and benefits of the options available to manage that risk, they will also need to consider the various constraints to action faced by the actors involved. Some of these decision-making contexts, such as the design of large infrastructure projects, may require rigorous quantitative information to feed formal evaluations, often including cost-benefit analysis (e.g. PriceWaterHouseCoopers, 2010; and see chapter 17). Others, especially at local level, such as decision-making in traditional communities, are often made more intuitively, with a much greater role for a wide range of social and cultural aspects. These may benefit much more from experience-based approaches, participatory risk assessments or story-telling to evaluate future implications of possible decisions (e.g. Van Aalst et al., 2008, World Bank, 2010). Multi-criteria analysis, scenario planning, and flexible decision paths offer options for taking action when faced with large uncertainties or incomplete information, and can help bridge adaptation strategies across scales (in particular between the national and local level). In most cases, an understanding of the context in which the risk plays out, and the alternative options that may be considered to manage it, are not an afterthought, but a defining feature of an appropriate climate risk analysis, which requires a much closer interplay between decision-makers and providers of climate risk information than often occurs in practice (e.g., Cardona et al., 2012; Hellmuth et al., 2010; Mendler de Suarez et al., 2012). The different decision-making contexts also determine the types of climate information required, including the climate variables of interest and the geographic and time scales on which they need to be provided. Many climate change impact assessments have traditionally focused on changes over longer time horizons (often out to 2100, though recently studies have begun to concentrate more on mid-century or earlier). In contrast, most decisions taken today have a planning horizon ranging from a few months to about two decades (e.g. Wilby et al., 2009). For many such shorter-term decisions, recent climate variability and observed trends are commonly regarded as sufficient to inform adaptation (e.g. Hallegatte, 2009). However, in so doing, there is often scope to make better use of observed climatological information as well as seasonal and maybe also decadal climate forecasts (e.g. Wang et al., 2009; Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ziervogel et al., 2010; Mehta et al., 2011; Kirtman et al., 2014., HLT, 2011). For longer-term decisions, such as decisions with irreversible long-term implications and investments with a long investment horizon and substantial vulnerability to changing climate conditions, longer-term climate risk information is needed (e.g. Reeder and Ranger, 2010). However, while that longer-term information is often used simply to plan for a best-guess scenario to optimize for most probable conditions, there is increasing attention for informing concerns about maladaptation (Barnett and O'Neill, 2010) and sequencing of potential adaptation options in a wider range of possible outcomes, requiring a stronger focus on ranges of possible outcomes and guidance on managing uncertainties, especially at regional, national and sub-national levels (Hall et al., 2012; Gersonius et al., 2013). Section 21.3 summarizes different approaches that have been applied at different scales looking at vulnerabilities and impacts, adaptation, and climate science in a regional context, paying special attention to information contained in the regional chapters. 21.2.2. Defining Regions There has been an evolution in the treatment of regional aspects of climate change in IPCC reports (Table 21-2) from a patchwork of case examples in the First Assessment Report (FAR) and its supplements, through to attempts at a more systematic coverage of regional issues following a request from governments, beginning with the Special Report on the Regional Impacts of Climate Change in 1998. That report distilled information from the Second Assessment Report (SAR) for ten continental scale regions, and the subsequent Third (TAR) and Fourth (AR4) assessments each contained comparable chapters on impacts, adaptation and vulnerability in the Working Group (WG) II volumes. WG I and III reports have also addressed regional issues in various chapters, using different methods of mapping, statistical aggregation and spatial averaging to provide regional information. [INSERT TABLE 21-2 HERE Table 21-2: Selected examples of regional treatment in previous IPCC Assessment Reports and Special Reports (SR). Major assessments are subdivided by the three Working Group reports, each described by generic titles.] Part B of this WG II Fifth Assessment (AR5) is the first to address regional issues treated in all three WGs. It comprises chapters on the six major continental land regions, Polar Regions, Small Islands and The Ocean. These are depicted in Figure 21-1. [INSERT FIGURE 21-1 HERE Figure 21-1: Specification of the world regions described in chapters 22-30 of this volume.] Some of the main topics benefiting from a regional treatment are: Changes in climate, typically represented over sub-continental regions, a scale at which global climate models simulate well the pattern of observed surface temperatures, though more modestly the pattern of precipitation (Flato et al., 2014). While maps are widely used to represent climatic patterns, regional aggregation of this (typically gridded) information is still required to summarise the processes and trends they depict. Examples, including information on climate extremes, are presented elsewhere in this chapter, with systematic coverage of all regions provided in supplementary material. Selected time series plots of temperature and precipitation change from an atlas of global and regional climate projections accompanying the WG I report (Collins et al., 2014b) can also be found in several regional chapters of this volume. In Figure 21-1, the sub-continental regions used for summarising climate information are overlaid on a map of the nine regions treated in Part B. Changes in other aspects of the climate system, such as the cryosphere, oceans, sea level, and atmospheric composition. A regional treatment of these phenomena is often extremely important to gauge real risks, for example when regional changes in land movements and local ocean currents counter or reinforce global sea level rise (Nicholls et al., 2013). Climate change impacts on natural resource sectors, such as agriculture, forestry, ecosystems, water resources and fisheries, and on human activities and infrastructure, often with regional treatment according Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 to biogeographical characteristics (e.g. biomes, climatic zones, physiographic features such as mountains, river basins, coastlines or deltas, or combinations of these). Adaptive capacity, which is a measure of society's ability to adjust to the potential impacts of climate change, sometimes characterised in relation to social vulnerability (Füssel, 2010), and sometimes represented in regional statistics through the use of socio-economic indicators. Emissions of greenhouse gases and aerosols and their cycling through the Earth system (Blanco et al., 2014; Ciais et al., 2014). Human responses to climate change through mitigation and adaptation, which can require both global and regional approaches (e.g. Stavins et al., 2014; Agrawala et al., 2014; Somanathan et al., 2014; and see chapters 14 to 16). Detailed examples of these elements will be referred to throughout this chapter and the regional chapters that follow. Some of the more important international political groupings that are pertinent to the climate change issue are described and catalogued in Supplementary material (section SM21.1). Table SM21-1 in section SM21.1 lists UN member states and other territories, their status in September 2013 with respect to some illustrative groupings of potential relevance for international climate change policy making, and the regional chapters in which they are considered in this report. Finally, new global socioeconomic and environmental scenarios for climate change research have emerged since the AR4 that are richer, more diverse and offer a higher level of regional detail than previous scenarios taken from the IPCC Special Report on Emissions Scenarios (SRES). These are introduced in Box 21-1. _____ START BOX 21-1 HERE _____ Box 21-1. A New Framework of Global Scenarios for Regional Assessment The major socio-economic driving factors of future emissions and their effects on the global climate system were characterized in the TAR and AR4 using scenarios derived from the IPCC Special Report on Emissions Scenarios (SRES IPCC, 2000). However, these scenarios are becoming outdated in terms of their data and projections, and their scope is too narrow to serve contemporary user needs (Ebi et al., 2013). More recently a new approach to developing climate and socio-economic scenarios has been adopted in which concentration trajectories for atmospheric greenhouse gases and aerosols were developed first (Representative Concentration Pathways or RCPs Moss et al., 2010), thereby allowing climate modeling work to proceed much earlier in the process than for SRES. Different possible Shared Socio-economic Pathways (SSPs), intended for shared use among different climate change research communities, were to be determined later, recognizing that more than one socio-economic pathway can lead to the same concentrations of greenhouse gases and aerosols (Kriegler et al., 2012). Four different RCPs were developed, corresponding to four different levels of radiative forcing of the atmosphere by 2100 relative to pre-industrial levels, expressed in units of Wm-2: RCP 8.5, 6.0, 4.5, and 2.6 (van Vuuren et al., 2012). These embrace the range of scenarios found in the literature, and all except RCP 8.5 also include explicit stabilization strategies, which were missing from the SRES set. An approximate mapping of the SRES scenarios onto the RCPs on the basis of a resemblance in radiative forcing by 2100 is presented in chapter 1 (this volume), pairing RCP 8.5 with SRES A2 and RCP 4.5 with B1 and noting that RCP 6.0 lies between B1 and B2. No SRES scenarios result in forcing as low as RCP 2.6, though mitigation scenarios developed from initial SRES trajectories have been applied in a few climate model experiments (e.g. the E1 scenario Johns et al., 2011). In addition, five SSPs have been proposed, representing a wide range of possible development pathways (van Vuuren et al., 2013). An inverse approach is applied, whereby the SSPs are constructed in terms of outcomes most relevant to IAV and mitigation analysis, depicted as challenges to mitigation and adaptation (Chapter 20, Figure 20- 3). Narrative storylines for the SSPs have been outlined and preliminary quantifications of the socio-economic variables are underway (O'Neill et al., 2013). Priority has been given to a set of basic SSPs with the minimum detail and comprehensiveness needed to provide inputs to IAV and integrated assessment models, primarily at global or large regional scales. Building on the basic SSPs, a second stage will construct extended SSPs, designed for finer- scale regional and sectoral applications (O'Neill et al., 2013). Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 An overall scenario architecture has been designed for integrating RCPs and SSPs (Ebi et al., 2013; van Vuuren et al., 2013), for considering mitigation and adaptation policies using Shared Policy Assumptions (SPAs Kriegler et al., 2013) and for providing relevant socio-economic information at the scales required for IAV analysis (van Ruijven et al., 2013). Additional information on these scenarios can be found in chapter 1 of this report (section 1.1.3) and elsewhere in the assessment (Blanco et al., 2014.; Collins et al., 2014a.; Kunreuther et al., 2014). However, due to the time lags that still exist between the generation of RCP-based climate change projections in CMIP5 (the Coupled Model Intercomparison Project Taylor et al., 2012) and the development of SSPs, few of the IAV studies assessed in this report actively use these scenarios. Instead, most of the scenario-related studies in the assessed literature still rely on the SRES. _____ END BOX 21-1 HERE _____ 21.2.3. Introduction to Methods and Information There has been significant confusion and debate about the definitions of key terms (Janssen and Ostrom, 2006), such as vulnerability (Adger, 2006), adaptation (Stafford Smith et al., 2011), adaptive capacity (Smit and Wandel, 2006) and resilience (Klein et al., 2003). One explanation is that the terms are not independent concepts, but defined by each other, thus making it impossible to remove the confusion around the definitions (Hinkel, 2011). The differences in the definitions relate to the different entry points for looking at climate change risk (IPCC, 2012). Table 21-3 shows two ways to think about vulnerability, demonstrating that different objectives (e.g., improving well-being and livelihoods or reducing climate change impacts) lead to different sets of questions being asked. This results in the selection of different methods to arrive at the answers. The two approaches portrayed in the middle and right hand columns of Table 21-3 have also been characterised in terms of top-down (middle column) and bottom-up (right column) perspectives, with the former identifying physical vulnerability and the latter social vulnerability (Dessai and Hulme, 2004). In the middle column, the climate change impacts are the starting point for the analysis, revealing that people and/or ecosystems are vulnerable to climate change. This approach commonly applies global- scale scenario information and seeks to refine this to the region of interest through downscaling procedures. For the approach illustrated on the right, the development context is the starting point (i.e., social vulnerability), commonly focusing on local scales, on top of which climate change occurs. The task is then to identify what changes are needed in the broader-scale development pathways in order to reduce vulnerability to climate change. Another difference is a contrast in time-frames, where a climate change focused approach tends to look to the future to see how to adjust to expected changes, whereas a vulnerability focused approach is centred on addressing the drivers of current vulnerability. A similar approach is described by McGray et al. (2009). [INSERT TABLE 21-3 HERE Table 21-3: Two possible entry points for thinking about vulnerability to climate change (illustrative and adapted from Füssel 2007).] The information assessed in this chapter stems from different entry points, framings and conceptual frameworks for thinking about risk. They merge social and natural science perspectives with transdisciplinary ones. There is no single "best" conceptual model: the approaches change as scientific thinking evolves. The IPCC itself is an example of this: IPCC SREX (IPCC, 2012) presented an approach that has been adjusted and adapted in Chapter 19 of this volume. Chapter 2 describes other conceptual models for decision making in the context of risk. While this diversity in approaches enriches our understanding of climate change, it can also create difficulties in comparisons. For instance, findings that are described as vulnerabilities in some studies may be classified as impacts in others; lack of adaptive capacity in one setting might be described as social vulnerability in another. Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 21.3. Synthesis of Key Regional Issues This section presents information on vulnerabilities and impacts, adaptation, and climate science in a regional context. To illustrate how these different elements play out in actual decision-making contexts, Table 21-4 presents examples drawn from the regional and thematic chapters, which illustrate how information about vulnerability and exposure, and climate science at different scales, inform adaptation (implemented in policy and practice as part of a wider decision-making context). These show that decision-making is informed by a combination of different types of information. However, this section is organized by the three constituent elements: vulnerabilities and impacts; adaptation; and climate science. [INSERT TABLE 21-4 HERE Table 21-4: Illustrative examples of adaptation experience, as well as approaches to reduce vulnerability and enhance resilience. Adaptation actions can be influenced by climate variability, extremes, and change, and by exposure and vulnerability at the scale of risk management. Many examples and case studies demonstrate complexity at the level of communities or specific regions within a country. It is at this spatial scale that complex interactions between vulnerabilities, inequalities, and climate change come to the fore. At the same time, place- based examples illustrate how larger-level drivers and stressors shape differential risks and livelihood trajectories, often mediated by institutions.] The following two sub-sections offer a brief synopsis of the approaches being reported in the different regional chapters on impacts and vulnerability studies (21.3.1) and adaptation studies (21.3.2), aiming to particularly highlight similarities and differences among regions. Table 21-5 serves as a rough template for organising this discussion which is limited to the literature that has been assessed by the regional chapters. It is organized according to the broad research approach applied, distinguishing impacts and vulnerability approaches from adaptation approaches; and according to scales of application ranging from global to local. [INSERT TABLE 21-5 HERE Table 21-5: Dimensions of assessments of impacts and vulnerability and of adaptation drawn upon to serve different target fields (cf. Table 21-1). Scales refer to the level of aggregation at which study results are presented. Entries are illustrations of different types of study approaches reported and evaluated in this volume, with references given both to the original studies and to the chapters in which they are cited. Aspects of some of the studies in this table are also alluded to in Section 21.5.] Section 21.3.3 then provides an analysis of advances in understanding of the physical climate system for the different regions covered in chapters 22-30, introducing new regional information to complement the large-scale and process-oriented findings presented by Working Group I. Understanding the reliability of this information is of crucial importance. In the context of IAV studies it is relevant to a very wide range of scales and it comes with a similarly wide range of reliabilities. Using a similar classification of spatial scales to that presented in Table 21-5, Table 21-6 provides a summary assessment of the reliability of information on two basic climate variables of relevance, surface temperature and precipitation. It is drawn from the extensive assessment and supporting literature from the IPCC SREX (IPCC, 2012) and the AR5 WG1 reports. Some discussion of relevant methodologies and related issues and results are also presented in section 21.5. [INSERT TABLE 21-6 HERE Table 21-6: Reliability of climate information on temperature and precipitation over a range of spatial and temporal scales. Reliability is assigned to one of seven broad categories from Very High (VH) to Medium (M) through to Very Low (VL).] Table 21-6 shows there are significant variations in reliability with finer scales information generally less reliable given the need for a greater density of observations and/or for models to maintain accuracy at high resolutions. The reliability of information on past climate depends on the availability and quality of observations which is higher for temperature than precipitation as observations of temperature are easier to make and generally more representative of surrounding areas than is the case for precipitation. Future climate change reliability depends on the performance of the models used for the projections in simulating the processes that lead to these changes. Again, information on Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 temperature is generally more reliable due to the models demonstrated ability to simulate the relevant processes when reproducing past changes. The significant geographical variations, in the case of the observations, result from issues with availability and/or quality of data in many regions, especially for precipitation. For future climate change, data availability is less of an issue with the advent of large ensembles of climate model projections but quality is a significant problem in some regions where the models perform poorly and there is little confidence that processes driving the projected changes are accurately captured. A framework for summary information on model projections of future climate change placed this in the context of observed changes is presented in Box 21-2. _____ START BOX 21-2 HERE _____ Box 21-2. Summary Regional Climate Projection Information Summary figures on observed and projected changes in temperature and precipitation are presented in the following regional chapters. These provide some context to the risks associated with climate change vulnerability and impacts and the decision-making on adaptations being planned and implemented in response to these risks. Figure 21-2 provides an example for Africa. The information is identical to that displayed in Box CC-RC. [INSERT FIGURE 21-2 HERE Figure 21-2: Observed and projected changes in annual average temperature and precipitation. (Top panel, left) Observed temperature trends from 1901-2012 determined by linear regression. [WGI AR5 Figures SPM.1 and 2.21] (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression. [WGI AR5 Figure SPM.2] For observed temperature and precipitation, trends have been calculated where sufficient data permits a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level. Diagonal lines indicate areas where change is not significant. (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and >90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in >66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See alsoAnnex I of WGI AR5. [Boxes 21-3 and CC-RC]] These figures provide a very broad overview of the projected regional climate changes but in dealing with only annual averages they are not able to convey any information about projected changes on seasonal timescales or shorter, such as for extremes. In addition, they are derived solely from the CMIP5 general circulation models (GCMs) and do not display any information derived from CMIP3 data which are widely used in many of the studies assessed within the AR5 WG2 report. To provide additional context two additional sets of figures are presented here and in Box 21-4 that display temperature and precipitation changes at the seasonal and daily timescales respectively. Figure 21-3 displays projected seasonal and annual changes averaged over the regions defined in the IPCC SREX report, Managing the risks of extreme events and disasters to advance climate change adaptation (IPCC 2012), for Central and South America for the four RCP scenarios and three of the SRES scenarios. The temperature and precipitation changes for the period 2071-2100 compared to a baseline of 1961-1990 are plotted for the four standard three months seasons with the changes from each CMIP3 or CMIP5 represented by a symbol. Symbols showing the CMIP3 model projections are all grey but differ in shape depending on the driving SRES concentrations scenario and those showing the CMIP5 projections differ in colour depending on the driving RCP emissions/concentrations scenario (see figure legend for details and Box 21-1 for more information on the SRES and RCP scenarios). The thirty year periods were chosen for consistency with the figures displayed in Box 21-4 (Figures 21-7 and 21-8) showing changes in daily temperatures and precipitation. Figures presenting similar information for Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 the SREX regions contained in the other inhabited continents are presented in supplementary figures SM21-1 to SM21-7. [INSERT FIGURE 21-3 HERE Figure 21-3: Regional average change in seasonal and annual mean temperature and precipitation over five sub- regions covering South and Central America for the period 2071-2100 relative to 1961-90 in GCM projections from 35 CMIP5 ensemble under four RCP scenarios (van Vuuren et al., 2011) compared with GCM projections from 22 CMIP3 ensemble under three SRES scenarios (IPCC, 2000); see Table 21-1 for details of the relationship between the SRES and RCP scenarios. Regional averages are based on SREX region definitions (see Figure 21-3). Temperature changes are given in C and precipitation changes in mm/day with axes scaled relative to the maximum changes projected across the range of models. The models which generated the data displayed are listed in supplementary material Table SM21-3.] _____ END BOX 21-2 HERE _____ 21.3.1. Vulnerabilities and Impacts 21.3.1.1. Observed Impacts The evidence linking observed impacts on biological, physical and (increasingly) human systems to recent and ongoing regional climate changes has become more compelling since the AR4 (see chapter 18). One reason for this is the improved reporting of published studies from hitherto under-represented regions of the world, especially in the tropics (Rosenzweig and Neofotis, 2013). That said, the disparity is still large between the copious evidence being presented from Europe and North America, as well as good quality data emerging from Australasia, polar regions, many ocean areas and some parts of Asia and South America, compared to the much sparser coverage of studies from Africa, large parts of Asia, central and South America and many small islands. On the other hand, as the time series of well-calibrated satellite observations become longer in duration, and hence statistically more robust, these are increasingly providing a near global coverage of changes in surface characteristics such as vegetation, hydrology, and snow and ice conditions that can usefully complement or substitute for surface observations (see Table 21-4 and chapter 18 for examples). Changes in climate variables other than temperature, such as precipitation. evapotranspiration and CO2 concentration, are also being related to observed impacts in a growing number of studies (Rosenzweig and Neofotis, 2013; and see examples from Australia in chapter 25, Table 25-3 and south eastern South America in chapter 27, Figure 27-7). Other regional differences in observed changes worth pointing out include trends in relative sea level, which is rising on average globally (Church et al., 2014), but displays large regional variations in magnitude, or even sign, due to a combination of influences ranging from El Nino/La Nina cycles to local tectonic activity (Nicholls et al., 2013), making general conclusions about ongoing and future risks of sea level change very difficult to draw across diverse regional groupings such as small islands (see chapter 29). There are also regional variations in another ongoing effect of rising CO2 concentration ocean acidification, with a greater pH decrease at high latitudes consistent with the generally lower buffer capacities of the high latitude oceans compared to lower latitudes (Rhein et al., 2014, section 3.8.2). Calcifying organisms are expected to show responses to these trends in future, but key uncertainties remain at organismal to ecosystem levels (chapter 30, Box CC-OA). 21.3.1.2. Future Impacts and Vulnerability 21.3.1.2.1. Impact models The long-term monitoring of environmental variables, as well as serving a critical role in the detection and attribution of observed impacts, also provides basic calibration material used for the development and testing of impact models. These include process-based or statistical models used to simulate the biophysical impacts of climate on outcomes such as crop yield, forest productivity, river runoff, coastal inundation or human mortality and Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 morbidity (see chapters 2-7; 11). They also encompass various types of economic models that can be applied to evaluate the costs incurred by biophysical impacts (see, for example, chapters 10 and 17). There are also integrated assessment models (IAMs), Earth System Models, and other more loosely linked integrated model frameworks that represent multiple systems and processes (e.g. energy, emissions, climate, land use change, biophysical impacts, economic effects, global trade) and the various interactions and feedbacks between them. For examples of these, see chapter 17, section 17.6.3 and Flato et al. (2014). 21.3.1.2.2. Vulnerability mapping A second approach to projecting potential future impacts is to construct vulnerability maps. These usually combine information on three components: exposure to a hazard (commonly defined by the magnitude of climate change, sensitivity to that hazard), the magnitude of response for a given level of climate change, and adaptive capacity (describing the social and economic means to withstand the impacts of climate change (IPCC, 2001)). Key indicators are selected to represent each of the three components, which are sometimes combined into a single index of vulnerability. Indicators are usually measured quantities taken from statistical sources (e.g. income, population), or have been modelled separately (e.g. key climate variables). Vulnerability indices have received close scrutiny in several recent reviews (Füssel, 2010; Hinkel, 2011; Malone and Engle, 2011; Preston et al., 2011; Kienberger et al., 2012), and a number of global studies have been critiqued by Füssel (2010). A variant of vulnerability mapping is risk mapping (e.g., Tran et al., 2009; Ogden et al., 2008). This commonly identifies a single indicator of hazard (e.g. a level of flood expected with a given return period), which can be mapped accurately to define those regions at risk from such an event (e.g. in a flood plain). Combined with information on changing return periods of such events under a changing climate would enable some estimate of altered risk to be determined. 21.3.1.2.3. Experiments A final approach for gaining insights on potential future impacts, concerns physical experiments designed to simulate future altered environments of climate (e.g. temperature, humidity and moisture), and atmospheric composition (e.g. CO2, surface ozone and sulphur dioxide concentrations ). These are typically conducted to study responses of crop plants, trees and natural vegetation, using open top chambers, greenhouses or free air gas release systems (e.g., Craufurd et al., 2013), or responses of aquatic organisms such as plankton, macrophytes or fish, using experimental water enclosures known as mesocosms (e.g., Sommer et al., 2007; Lassen et al., 2010). 21.3.1.2.4. Scale issues Impact models operate at a range of spatial and temporal resolutions, and while their outputs are sometimes presented as fine resolution maps, key model findings are rarely produced at the finest resolution of the simulations (i.e. they are commonly aggregated to political or topographic units of interest to the target audience, e.g. watershed, municipality, national or even global). Aggregation of data to coarse-scale units is also essential for allowing comparison of outputs from models operating at different resolutions, but it also means that sometimes quite useful detail may be overlooked when model outputs are presented at the scale of the coarsest common denominator. Conversely, if outputs from impact models are required as inputs to other models, the outputs may need to be harmonized to a finer grid than the original data. In such cases, downscaling methods are commonly applied. This was the case, for example, when providing spatially explicit projections of future land use from different IAMs (Hurtt et al., 2011) for climate modellers to apply in the CMIP5 process (Collins et al., 2014a). It is also a common procedure used in matching climate model outputs to impact models designed to be applied locally (e.g. over a river basin or an urban area see section 21.3.3.2). Even if the same metrics are being used to compare aggregate model results (e.g. developed versus developing country income under a given future scenario) estimates may have been obtained using completely different types of Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 models operating at different resolutions. Moreover, many models that have a large-scale coverage (e.g. continental or global) may nonetheless simulate processes at a relatively fine spatial resolution, offering a potentially useful source of spatially explicit information that is unfamiliar to analysts working in specific regions, who may defer to models more commonly applied at the regional scale. Examples include comparison of hydrological models with a global and regional scope (Todd et al., 2011) and bioclimatic models of vascular plant distributions with a European and local scope (Trivedi et al., 2008). Vulnerability mapping exercises can also be undermined by the inappropriate merging of indicator datasets that resolve information to a different level of precision (e.g. Tzanopoulos et al., 2013). There is scope for considerably enhanced cross-scale model intercomparison work in the future, and projects such as AgMIP (Rosenzweig et al., 2013) and ISI-MIP (Schiermeier, 2012, see section 21.5) have provision for just such exercises. 21.3.2. Adaptation This section draws on material from the regional chapters (22-30) as well as the examples described in Table 21-4. Material from chapters 14-17 is also considered. See also Table 16.4 for a synthesis from the perspective of adaptation constrains and limits. 21.3.2.1. Similarities and Differences in Regions As described in the regional chapters, a large portion of adaptation knowledge is based on conclusions drawn from case studies in specific locations, the conceptual findings are typically being applied globally (chapters 14-17). ). It is this empirical knowledge on adaptation that guides understandings in the different regions. This is especially the case for developing regions. Thus, regional approaches to adaptation vary in their degree of generality. One of the most striking differences between regions in terms of adaptation is the extent to which it has been studied and implemented. Australia and Europe have invested heavily in research on adaptation since the AR4, and the result is a rich body of literature published by local scientists. The ability to advance in adaptation knowledge may be related to the amount and quality of reliable climate information, the lack of which has been identified as a constraint to developing adaptation measures in Africa (22.4.2). Many case studies, especially of community-based adaptation, stem from Asia, Africa, Central and South America and Small Islands but the majority of this work has been undertaken and authored by international non-governmental organisations, as well as by other non-local researchers. In Africa, most planned adaptation work is considered to be pilot, and seen as part of learning about adaptation, although there has been significant progress since the AR4 (22.4.4.2). Most regional chapters report lags in policy work on adaptation (see also 16.5.2). While most European countries have adaptation strategies, few have been implemented (23.1.2). Lack of implementation of plans is also the case for Africa (22.4). In North (26.8.4.1.2) and Central and South America (27.5.3.2), adaptation plans are in place for some cities. In Australasia, there are few adaptation plans (25.4.2). In the Arctic, they are in their infancy (28.4). At the same time, civil society and local communities have the opportunity to play a role in decision making about adaptation in Europe and Asia (23.7.2, 24.4.6.5). In Africa, social learning and collective action are used to promote adaptation (22.4.5.3). Adaptation is observed as mostly autonomous (spontaneous) in Africa, although socio- ecological changes are creating constraints for autonomous adaptation (22.4.5.4). There is a disconnect in most parts of Africa between policy and planning levels, and the majority of work is still autonomous and unsupported (22.4.1). In the case of UNFCCC-supported activities, such as National Adaptation Programmes of Action, few projects from the African (22.4.4.2) least developed countries have been funded, thus limiting the effectiveness of these investments. Several chapters (Africa, Europe, North America, Central and South America and Small Islands) explicitly point out that climate change is only one of multiple factors that affect societies and ecosystems and drives vulnerability or challenges adaptation (22.4.2, 23.10.1, 26.8.3.1, 27.3.1.2, 29.6.3). For example, North America reports that for water resources, most adaptation actions are no-regrets , meaning that they have benefits beyond just adaptation to climate change (26.3.4). In Australasia, the limited role of socio-economic information in vulnerability assessments restricts confidence regarding the conclusions about future vulnerability and adaptive capacity (25.3.2). Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Some chapters (Polar Regions, North America, Australasia) emphasise the challenges faced by indigenous peoples and communities in dealing with climate change (28.4.1, 26.8.2.2, 25.8.2). Although they are described as having some degree of adaptive capacity to deal with climate variability, shifts in lifestyles combined with a loss of traditional knowledge leave many groups more vulnerable to climate change (28.2.4.2). Also, traditional responses have been found to be maladaptive because they are unable to adjust to the rate of change, or the broader context in which the change is taking place, as seen in the Arctic (28.4.1). In response to changing environmental conditions, people are taking on maladaptive behaviour for instance, by going further to hunt because of changed fish-stocks and thus exposing themselves to greater risk, or changing to different species and depleting stocks (28.4.1). Limits to traditional approaches for responding to changing conditions have also been observed in several Small Island States (29.8). Most populated regions have experience with adaptation strategies in agriculture, where exposure to the impacts of climate variability over centuries provides a starting point for making adjustments to new changes in climate. Water and land use management strategies stand out in the literature in common across all of the main continental regions. The link between adaptation and development is explicit in Africa, where livelihood diversification has been key to reducing vulnerability (22.4.5.2). At the same time, there is evidence that many short-term development initiatives have been responsible for increasing vulnerability (22.4.4.2). Other chapters mention constraints or barriers to adaptation in their regions. For example, the low priority accorded to adaptation in parts of Asia, compared to more pressing issues of employment and education, is attributed in part to a lack of awareness of the potential impacts of climate change and the need to adapt, a feature common to many regions (22.5.4). All developing regions cite insufficient financial resources for implementing adaptation as a significant limitation. 21.3.2.2. Adaptation Examples in Multiple Regions Across regions, similar responses to climate variability and change can be noted. Heat waves are an interesting example (Table 21-4), as early warning systems are gaining use for helping people reduce exposure to heat waves. At the global scale, the length and frequency of warm spells, including heat waves, has increased since 1950 (medium confidence), and over most land areas on a regional scale, more frequent and/or longer heat waves or warm spells are likely by 2016-2035 and very likely by 2081-2100 (IPCC, 2014d). Warning systems are now planned and implemented in Europe, the US, Canada, Asia and Australia. Use of mangroves to reduce flood risks and protect coastal areas from storm surges is a measure promoted in Asia, Africa, the Pacific and South America (Table 21-4). Often, mangroves have been cut down to provide coastal access, so there is a need to restore and rehabilitate them. This is an example that is considered low-regrets because it brings multiple benefits to communities besides protecting them from storm surges, such as providing food security and enhancing ecosystem services. Mangrove forests also store carbon, offering synergies with mitigation. In several African countries, as well as in India, index-based insurance for agriculture has been used to address food insecurity and loss of crops resulting from more hot and fewer cold nights, an increase in heavy precipitation events and longer warm spells (Table 21-4). A predetermined weather threshold typically associated with high loss triggers an insurance pay-out. The mechanism shares risk across communities and can help encourage adaptive responses and foster risk awareness and risk reduction. However, limited availability of accurate weather data mean that establishing which weather conditions causes losses can be challenging. Furthermore, if there are losses but not enough to trigger payout, farmers may lose trust in the mechanism. 21.3.2.3. Adaptation Examples in Single Regions Although conditions are distinct in each region and location, practical lessons can often be drawn from looking at examples of adaptation in different locations. Experience with similar approaches in different regions offers additional lessons that can be useful when deciding whether an approach is appropriate. Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Community-based adaptation is happening and being planned in many developing regions, especially in locations that are particularly poor. In small islands, where a significant increase in the occurrence of future sea-level extremes by 2050 and 2100 is anticipated, traditional technologies and skills may still be relevant for adapting (Table 21-4). In the Solomon Islands, relevant traditional practices include elevating concrete floors to keep them dry during heavy precipitation events and building low aerodynamic houses with palm leaves as roofing to avoid hazards from flying debris during cyclones, supported by perceptions that traditional construction methods are more resilient to extreme weather. In Fiji after cyclone Ami in 2003, mutual support and risk sharing formed a central pillar for community-based adaptation, with unaffected households fishing to support those with damaged homes. Participatory consultations across stakeholders and sectors within communities and capacity building taking into account traditional practices can be vital to the success of adaptation initiatives in island communities, such as in Fiji or Samoa. These actions provide more than just the immediate benefits; they empower people to feel in control of their situations. In Europe, several governments have made ambitious efforts to address risks of inland and coastal flooding due to higher precipitation and sea level rise during the coming century (Table 21-4). Efforts include a multitude of options. One of the key ingredients is strong political leadership or government champions. In the Netherlands, government recommendations include soft measures preserving land from development to accommodate increased river inundation; raising the level of lakes to ensure continuous freshwater supply; restoring natural estuary and tidal regimes; maintaining coastal protection through beach nourishment; and ensuring necessary political-administrative, legal, and financial resources. The British government has also developed extensive adaptation plans to adjust and improve flood defenses and restrict development in flood risk areas in order to protect London from future storm surges and river flooding. They undertook a multi-stage process, engaging stakeholders and using multi-criteria analysis. Pathways have been analyzed for different adaptation options and decisions, depending on eventual sea level rise, with ongoing monitoring of the drivers of risk informing decisions. In Australia, farmers and industries are responding to experienced and expected changes in temperature, rainfall, and water availability by relocating parts of their operations, such as for rice, wine, or peanuts, or changing land use completely (Table 21-4). In South Australia, for instance, there has been some switching from grazing to cropping. The response is transformational adaptation, and can have positive or negative implications for communities in both origin and destination regions. This type of adaptation requires a greater level of commitment, access to more resources and greater integration across decision-making levels because it spans regions, livelihoods and economic sectors. 21.3.3. Climate System This section places the regional chapters in a broader context of regional climate information, particularly regarding cross regional aspects, but does not provide a detailed region-by-region assessment. Boxes 21-2 and 21-4 introduce examples of regional information for continental/sub-continental regions but other regional definitions are often relevant (see Box 21-3). The focus in this section is on the summary of new and emerging knowledge since the AR4 relevant to vulnerability, impacts and adaptation research, with emphasis on material deriving from dynamical and statistical downscaling work which is often of greater relevance for VIA applications than the coarser resolution global climate model data. In a regional context, the AR5 WG1 Chapter 14 (Regional Phenomena) is particularly relevant for the projections and evaluation of confidence in models ability to simulate temperature, precipitation and phenomena, together with an assessed implication for the general level of confidence in projections for 2080- 2099 of regional temperature and precipitation (See WG1, Ch 14, Table 14.2). _____ START BOX 21-3 HERE _____ Box 21-3. Developing Regional Climate Information Relevant to Political and Economic Regions In many world regions, countries form political and/or economic groupings that coordinate activities to further the interests of the constituent nations and their peoples. For example, the Intergovernmental Authority on Development (IGAD) of the countries of the Greater Horn of Africa recognizes that the region is prone to extreme climate events Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 such as droughts and floods that have severe negative impacts on key socio-economic sectors in all its countries. In response it has set up the IGAD Climate Prediction and Applications Centre (ICPAC). ICPAC provides and supports application of early warning and related climate information for the management of climate-related risks (for more details see http://www.icpac.net/). In addition it coordinates the development and dissemination of seasonal climate forecasts for the IGAD countries as part of a WMO-sponsored Regional Climate Outlook Forum process (Ogallo et al., 2008) which perform the same function in many regions. A more recent WMO initiative, the Global Framework for Climate Services (Hewitt et al., 2012), aims to build on these and other global, regional and national activities and institutions to develop climate information services for all nations. As socio-economic factors are important contributors to both the vulnerability and adaptability of human and natural systems, it clearly makes sense to summarise and assess available climate and climate change information for these regions, as this will be relevant to policy decisions taken within these groupings on their responses to climate change. For example, Figure 22-2 in Chapter 22 illustrates the presentation of observed and projected climate changes of two summary statistics for 5 political/economic regions covering much of Africa. It conveys several important pieces of information: the models are able to reproduce the observed trends in temperature; they simulate significantly lower temperatures without the anthropogenic forcings and significantly higher future temperatures under typical emissions paths; for most regions the models project that future variations in the annual average will be similar to those simulated for the past. However, for a more comprehensive understanding additional information needs to be included on other important aspects of climate, e.g. extremes (see Box 21-4). _____ END BOX 21-3 HERE _____ 21.3.3.1. Global context 21.3.3.1.1. Observed changes Temperature and precipitation New estimates of global surface air temperatures give a warming of about 0.89 C (0.69 C 1.08 C) for the period of 1901-2012 and about 0.72 C (0.49 C 0.79 C) for the period 1951-2012 (WGI Chapter 2, Section 2.4.3). Positive annual temperature trends are found over most land areas, particularly since 1981. Over the period 1981- 2012, relatively large trends have occurred over areas of Europe, the Sahara and middle East, central and northern Asia, north-eastern North America (WGI Chapter 2, Section 2.4.3). For precipitation, the Northern Hemisphere mid to high latitudes show a likely increasing trend (medium confidence prior to 1950, high confidence afterwards) (WGI Chapter 2, Section 2.5.1). Observed precipitation trends show a high degree of spatial and temporal variability, with both positive and negative values (WGI Chapter 2, Section 2.5). The human influence on warming since the middle of the 20th century is likely over every continental region, except Antarctica (WGI Chapter 10, Section 10.3.1), while the attribution of changes in hydrological variables is less confident (WGI Chapter 10, Section 10.3.2). Cryosphere New data have become available since the AR4 to evaluate changes in the cryosphere (WGI Chapter 4, Section 4.1) showing that the retreat of annual Arctic sea ice extent has continued, at a very likely rate of 3.5-4.1% per decade during the period 1979-2012. The perennial sea ice extent (sea ice area at summer minimum) decreased at a rate of 11.5 +/- 2.1% per decade (very likely) over the same period 1979-2012 (WGI Chapter 4, Section 4.2.2). The thickness, concentration and volume of arctic ice have also decreased. Conversely, the total annual extent of Antarctic ice has increased slightly (very likely 1.2-1.8% per decade between 1979 and 2011), with strong regional differences (WGI Chapter 4, Section 4.2.3). Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Almost all glaciers worldwide have continued to shrink since the AR4, with varying rates across regions (WG1 Chapter 4, Sections 4.3.1, 4.3.3). In particular, during the last decade most ice loss has been observed from glaciers in Alaska, the Canadian Arctic, the Southern Andes, the Asian mountains and the periphery of the Greenland Ice Sheet. Several hundred glaciers globally have completely disappeared in the last 30 years (WG1 Chapter 4, 4.3.3). Because of better techniques and more data, confidence has increased in the measurements of Greenland and Antarctica ice sheets. These indicate that parts of the Antarctic and Greenland ice sheets have been losing mass over the last two decades (high confidence), mostly due to changes in ice flow in Antarctica, and a mix of changes in ice flow and increases in snow/ice melt in Greenland. Ice shelves in the Antarctic Peninsula are continuing a long-term trend of thinning and partial collapse started some decades ago (WGI Chapter 4, Sections 4.4.2, 4.4.3, 4.4.5). 21.3.3.1.2. Near-term and long-term climate projections The uncertainty in near term CMIP5 projections is dominated by internal variability of the climate system (see Glossary entry on Climate Variability), initial ocean conditions and inter-model response, particularly at smaller spatial and temporal scales (Hawkins and Sutton 2009, 2011). In the medium and long term, emission profiles may affect the climate response. Global warming of 0.3-0.7C is likely for the period of 2016-2035 compared to 1986- 2005 based on the CMIP5 multi model ensemble, and spatial patterns of near term warming are generally consistent with the AR4 (WGI Chapter 11, Section 11.3.6). For precipitation (2016-2035 vs. 1986-2005), zonal mean precipitation will very likely increase in high and some of the mid-latitudes, and will more likely than not decrease in the subtropics (WGI Chapter 11, Section 11.3.2). Results from multi-decadal near term prediction experiments (up to 2035) with initialized ocean state show that there is some evidence of predictability of yearly to decadal temperature averages both globally and for some geographical regions (WGI Chapter 11, Section 11.2.3). Moving to long term projections (up to 2100), analyses of the CMIP5 ensemble have shown that, in general, the mean temperature and precipitation regional change patterns are similar to those found for CMIP3, with a pattern correlation between CMIP5 and CMIP3 ensemble mean late 21st century change greater than 0.9 for temperature and greater than 0.8 for precipitation (WGI Chapter 12, Section 12.4). Given the increased comprehensiveness and higher resolution of the CMIP5 models this adds an element of robustness to the projected regional change patterns. Some of the main characteristics of the projected late 21st century regional temperature and precipitation changes derived from the CMIP5 ensemble can be broadly summarized as follows (from WGI Chapter 12 and the WGI Atlas) with further details provided in Box 21-2 and accompanying supplementary material. Temperature Regions that exhibit relatively high projected temperature changes (often greater than the global mean by 50% or more) are high latitude Northern Hemisphere land areas and the Arctic, especially in December-January-February, and Central North America, portions of the Amazon, the Mediterranean, and Central Asia in June-July-August (Figure 21-4). [INSERT FIGURE 21-4 HERE Figure 21-4: CMIP5 ensemble median ratio of local:global average temperature change in the period 2071-2100 relative to 1961-90 under the RCP8.5 emissions/concentrations scenario. The values are displayed on a common 2.5x3.75 grid onto which each models data were regridded and they were calculated as follows: 1) for each model the local change was calculated between 1961 and 1990 at each grid cell, and is divided by the global average change in that model projection over the same period; 2) the median ratio value across all models at each grid cell is identified and shown. Data used are from the 35 CMIP5 models for which monthly projections were available under RCP8.5 which are listed in supplementary Table 21-3. Overplotted polygons indicate the SREX regions (IPCC, 2012) used to define the sub-regions used to summarise information in Chapters 21 and some of the subsequent regional chapters.] Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Precipitation Changes in precipitation are regionally highly variable, with different areas projected to experience positive or negative changes (Box 21-2). By the end of the century in the RCP8.5 scenario, the high latitudes will very likely experience greater amounts of precipitation, some mid-latitude arid and semi-arid regions will likely experience drying, while some moist mid-latitude regions will likely experience increased precipitation (WG1 Chapter 12, Section 12.4.5). Studies have also attempted to obtain regional information based on pattern scaling techniques in which regional temperature and precipitation changes are derived as a function of global temperature change (e.g. Giorgi 2008; Watterson 2008; Watterson and Whetton 2011; Watterson 2011 Ishizaki et al. 2012, 2013a, 2013b). Figure 21-5 from Harris et al. (2013) provides an example of Probability Density Functions (PDFs) of temperature and precipitation change over sub-continental scale regions obtained using a Bayesian method complemented by pattern scaling and performance-based model weighting. [INSERT FIGURE 21-5 HERE Figure 21-5: Evolution of the 5%, 17%, 33%, 50%, 66%, 83% and 95% percentiles of the distribution functions for annual surface air temperature changes (upper panel) and annual precipitation (lower panel) for the Giorgi-Francisco (2000) regions and the globe with the SRES A1B forcing scenario combining results from a perturbed physics ensemble and the CMIP3 ensemble. Twenty year means relative to the 1961-1990 baseline are plotted in decadal steps using a common y-axis scale. The 5%, 50% and 95% percentile values for the period 2080-2099 are displayed for each region. (From Harris et al. 2012).] 21.3.3.2. Dynamically and Statistically Downscaled Climate Projections Dynamical and statistical downscaling techniques have been increasingly applied to produce regional climate change projections, often as part of multi-model intercomparison projects (Görgen et al, 2010). A large number of RCM-based climate projections for the European region were produced as part of the European projects PRUDENCE (Christensen et al. 2007; Deque et al. 2007) and ENSEMBLES (Hewitt 2005; Deque and Somot 2010). High resolution projections (grid interval of ~12 km) were also produced as part of Euro-CORDEX (Jacob et al 2013). All these studies provide a generally consistent picture of seasonally and latitudinally varying patterns of change, which Giorgi and Coppola (2007) summarized with the term European Climate Change Oscillation (ECO) . The ECO consists of a dipole pattern of precipitation change, with decreased precipitation to the south (Mediterranean) and increased to the north (Northern Europe) following a latitudinal/seasonal oscillation. As a result, the Mediterranean region is projected to be much drier and hotter than today in the warm seasons (Giorgi and Lionello 2008), and central/northern Europe much warmer and wetter in the cold seasons (Kjellstrom and Ruosteenoja, 2007). An increase of interannual variability of precipitation and summer temperature is also projected throughout Europe, with a decrease in winter temperature variability over Northern Europe (Schar et al. 2004; Giorgi and Coppola 2007; Lenderink et al. 2007). This leads to broader seasonal anomaly distributions and a higher frequency and intensity of extreme hot and dry summers (e.g. Schar et al. 2004; Seneviratne et al. 2006; Beniston et al. 2007; Coppola and Giorgi 2010), for which a substantial contribution is given by land-atmosphere feedbacks (Seneviratne et al. 2006; Fischer et al. 2007; Seneviratne et al. 2010; Hirschi et al. 2011; Jaeger and Seneviratne 2011). The broad patterns of change in regional model simulations generally follow those of the driving global models (Christensen and Christensen 2007; Deque et al. 2007; Zanis et al. 2009), however fine scale differences related to local topographical, land use and coastline features are produced (e.g. Gao et al. 2006; Coppola and Giorgi 2010; Tolika et al. 2012). As part of the ENSEMBLES and AMMA projects, multiple RCMs were run for the period 1990-2050 (A1B scenario) over domains encompassing the West Africa region with lateral boundary conditions from different GCMs. The RCM-simulated West Africa monsoon showed a wide range of response in the projections, even when the models were driven by the same GCMs (Paeth et al. 2011) (Figure 21-6). Although at least some of the response patterns may be within the natural variability, this result suggests that for Africa, and probably more generally the Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 tropical regions, local processes and how they are represented in models play a key factor in determining the precipitation change signal, leading to a relatively high uncertainty (Engelbrecht et al. 2009; Haensler et al. 2011; Mariotti et al. 2011; Diallo et al. 2012). Statistical downscaling techniques have also been applied to the Africa region (Hewitson and Crane 2006; Lumsden et al. 2009; Paeth and Diederich 2010; Goergen et al., 2010; Benestad 2011). In this regard, methodological developments since the AR4 have been limited (see, for example reviews in Brown et al. 2008; Paeth et al. 2011) and activities have focused more on the applications (e.g. Mukheibir 2007; Gerbaux et al. 2009) for regional specific activities in the context of IAV work. [INSERT FIGURE 21-6 HERE Figure 21-6: Linear changes (i.e. changes obtained by fitting the time series at each grid point with straight lines) of annual precipitation during the 2001-2050 period from 10 individual RCM experiments and the MME mean under the A1B emission scenario. The top middle panels also account for projected land cover changes (see Paeth et al. 2011 for further explanation). Note that the REMO trends in both panels arise from a three-member ensemble whereas all other RCMs are represented by one single simulation. Trends statistically significant at the 95% level are marked by black dots. From Paeth et al. (2011).] Several RCM and time-slice high resolution GCM experiments have been conducted or analyzed for the South America continent (Nunez et al. 2009; Menendez et al. 2010; Sorensson et al. 2010; Marengo et al. 2009, 2010; Cabre et al. 2010; Kitoh et al. 2011). Overall, these studies revealed varied patterns of temperature and precipitation change, depending on the global and regional models used, however a consistent change found in many of these studies was an increase in both precipitation intensity and extremes, especially in areas where mean precipitation was projected to also increase. The Central American region has emerged as a prominent climate change hot-spot since the AR4, especially in terms of a consistent decrease of precipitation projected by most models, particularly in June-July (Rauscher et al 2008, 2011). Regional model studies focusing specifically on Central America projections are however still too sparse to provide robust conclusions (e.g. Campbell et al. 2010). Since the AR4 there has been considerable attention to producing higher resolution climate change projections over North America based on RCMs and high resolution global time slices (e.g. Salathe et al. 2008, 2010; Dominguez et al. 2010; Subin et al. 2011), in particular as part of the North American Regional Climate Change Assessment Program (NARCCAP Mearns et al. 2009; 2012; 2013). Results indicate variations (and thus uncertainty) in future climate based on the different RCMs, even when driven by the same GCM in certain subdomains (De Elia and Cote 2010; Mearns et al., 2013; Bukovsky et al. 2013). However, in the NARCCAP suite of simulations there were also some important commonalities in the climate changes produced by the RCMs. For example, they produced larger and more consistent decreases in precipitation throughout the Great Plains in summer than did the driving GCMs or the full suite of CMIP3 GCM simulations as well as larger increases in precipitation in the northern part of the domain in winter. In the realm of statistical downscaling and spatial disaggregation, considerable efforts have been devoted to applying different statistical models for the entire US and parts of Canada (e.g., Maurer et al. 2007; Hayhoe et al. 2010; Schoof et al., 2010). Numerous high resolution RCM projections have been carried out over the East Asia continent. While some of these find increases in monsoon precipitation over South Asia in agreement with the driving GCMs (Kumar et al. 2013) others also produce results that are not in line with those from GCMs. For example, both Ashfaq et al. (2009) and Gao et al. (2011) found in high resolution RCM experiments (20 and 25 km grid spacing, respectively) decreases in monsoon precipitation over areas of India and China in which the driving GCMs projected an increase in monsoon rain. Other high resolution (20 km grid spacing) projections include a series of double nested RCM scenario runs for the Korea peninsula (Im et al. 2007a; 2008a,b; 2010; 2011; Im and Ahn, 2011) indicating a complex fine scale structure of the climate change signal in response to local topographical forcing. Finally, very high resolution simulations were also performed. Using a 5-km mesh non-hydrostatic RCM nested within a 20-km mesh AGCM, Kitoh et al. (2009) and Kanada et al. (2012) projected a significant increase in intense daily precipitation around western Japan during the late Baiu season. Finally, a range of RCM, variable resolution and statistical downscaling 21st century projections have been conducted over the Australian continent or some of its sub-regions (Nunez and Mc Gregor 2007; Watterson et al. Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2008; Song et al. 2008, Timbal et al. 2008; Yin et al. 2010, Grose et al 2012a,b; Bennett et al 2012), showing that a local fine scale modulation of the large scale climate signal occurs in response to topographical and coastal forcings. 21.3.3.3. Projected Changes in Hydroclimatic Regimes, Major Modes of Variability, and Regional Circulations By modifying the Earth s energy and water budgets, climate change may possibly lead to significant changes in hydroclimatic regimes and major modes of climate variability (Trenberth et al. 2003). For example, Giorgi et al. (2011) defined an index of hydroclimatic intensity (HY-INT) incorporating a combined measure of precipitation intensity and mean dry spell length. Based on an analysis of observations, global and regional climate model simulations, they found that a ubiquitous global and regional increase in HY-INT was a strong hydroclimatic signature in model projections consistent with observations for the late decades of the 20th century. This suggests that global warming may lead to a hydroclimatic regime shift towards more intense and less frequent precipitation events, which would increase the risk of both flood and drought associated with global warming. ENSO is a regional mode of variability that substantially affects human and natural systems (Mc Phaden et al. 2006). Although model projections indicate that ENSO remains a major mode of tropical variability in the future, there is little evidence to indicate changes forced by GHG warming which are outside the natural modulation of ENSO occurrences (WGI Chapter 14, Sections 14.4, 14.8). The North Atlantic Oscillation (NAO) is a major mode of variability for the northern Hemisphere mid-latitude climate. Model projections indicate that the NAO phase is likely to become slightly more positive (WGI Chapter 14 ES) due to GHG forcing, but the NAO will be dominated by its large natural fluctuations. Model projections indicate that the Southern Annular Mode (SAM), a major mode of variability for the southern hemisphere, is likely going to weaken as ozone concentrations recover through the mid-21st century (WGI Chapter 14, Sections 14.5, 14.8). Regional circulations, such as the monsoon, are expected to change. The global monsoon precipitation, aggregated over all monsoon systems, is likely to strengthen in the 21st century with increases in its area and intensity, while the monsoon circulation weakens. Different regional monsoon systems, however, exhibit different responses to GHG forcing in the 21st century (WGI Chapter 14, Section 14.2.1). 21.3.3.4. Projected Changes in Extreme Climate Events CMIP5 projections confirm results from the CMIP3; a decrease in the frequency of cold days and nights, an increase in the frequency of warm days and nights, an increase in the duration of heat waves and an increase in the frequency and intensity of high precipitation events, both in the near term and far future (IPCC (2012), 3.3.2, 3.4.4; WGI Chapter 12, 12.4.5). Increases in intensity of precipitation (and thus risk of flood) and summer drought occurrence over some mid-continental land areas is a robust signature of global warming, both in observations for recent decades and in model projections (Trenberth 2011; WG1 Chapter 12, 12.4.5). For tropical cyclones there is still little confidence in past trends and near term projections (Seneviratne et al 2012). Globally, tropical cyclone frequency is projected to either not change or decrease and, overall, wind speed and precipitation is likely to increase though basin scale specific conclusions are still unclear (Knutson et al. 2010). A summary of observed and projections extremes along with some statistics on CMIP5 projections of changes in daily temperature and precipitation extremes over the main continents and the SREX regions (Figure 21-4) are introduced in Box 21-4 and accompanying supplementary material. _____ START BOX 21-4 HERE _____ Box 21-4. Synthesis of Projected Changes in Extremes Related to Temperature and Precipitation The IPCC report, Managing the risks of extreme events and disasters to advance climate change adaptation (IPCC 2012), or SREX for short, provides an in depth assessment of observed and projected changes in climate extremes. Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Due to the relevance of this material for assessing risks associated with climate change vulnerability and impacts and responses to these risks, summary information is presented here both drawing from and building on the material in the SREX report, including additional analyses of CMIP5 data (only CMIP3 data were used in SREX). Summaries of SREX findings relevant to three continents Latin America (Cameron et al., 2012), Asia and Africa (available from http://cdkn.org/srex/) have been developed using material from SREX Chapter 3. A synthesis of this material for all SREX regions, along with additional material from WG1 AR5, is presented in Table 21-7. This demonstrates that in many areas of the world there is higher confidence in future changes in extreme events than there is in past trends, often due to a lack of evidence on observed changes. [INSERT TABLE 21-7 HERE Table 21-7: An assessment of observed and projected future changes in temperature and precipitation extremes over 26 sub-continental regions as defined in the SREX report (IPCC 2012; see also Figure 21-4 and Table SM21.2). Confidence levels are indicated by colour coding of the symbols. Likelihood terms are given only for high confidence statements and are specified in the text. Observed trends in temperature and precipitation extremes, including dryness, are generally calculated from 1950, using the period 1961-1990 as a baseline (see Box 3-1 of IPCC (2012a)). The future changes are derived from global and regional climate model projections of the climate of 2071-2100 compared with 1961-1990 or 2080-2100 compared with 1980-2000. Table entries are summaries of information in Tables 3-2 and 3-3 of IPCC (2012a) supplemented with or superseded by material from Chapters 2 (section 2.6 and Table 2-13) and 14 (section 14.4) of the IPCC AR5 WG1 report and Table 25-1 of the IPCC WG2 report. The source(s) of information for each entry are indicated by the superscripts a (Table 3-2 of IPCC, 2012a), b (Table 3-3 of IPCC, 2012a), c (Chapter 2 (section 2.6 and Table 2-13) IPCC AR5 WG1 report), d (Chapter 14 (section 14.4) of the IPCC AR5 WG1 report) and e (Table 25-1 of the IPCC WG2 report).] In the SREX report, the only coordinated global multi-model ensemble information available was from the CMIP3. In order to provide information consistent with the projections assessed elsewhere in WG1 and 2, changes in daily temperature and precipitation projected by the CMIP5 models are presented here for two example indices, the 90th percentiles of the daily maximum temperature and daily precipitation amounts on wet days. Changes in these indices were calculated over 30 year periods (1961-1990 for the baseline and two future periods, 2041-2070 and 2071-2100) and the analysis was focused on the less extreme daily events to reduce problems with the number needed to be sampled to generate robust statistics (Kendon et. al, 2008). Projected changes were calculated for RCPs 4.5 and 8.5 and the results are displayed as a map for a given continental region and also regional averages over the SREX regions within that continent. Two examples are provided, for temperature changes over N America (Figure 21-7) and precipitation changes over Asia (Figure 21-8). A full set can be found in supplementary Figures SM21-8 to SM21-19. [INSERT FIGURE 21-7 HERE Figure 21-7: The frequency of 'warm days' (defined here as the 90th percentile daily maximum temperature during a baseline period of 1961-1990) projected for the 2071-2100 period by 26 CMIP5 GCMs for North America. Map: Ensemble median frequency of 'warm days' during 2071-2100 under RCP8.5. Graphs: Box-and-whisker plots indicate the range of regionally-averaged 'hot-day' frequency by 2041-2070 and 2071-2100 under RCPs 4.5 and 8.5 across the 26 CMIP5 models for each SREX sub-regions in North America. Boxes represent inter-quartile range and whiskers indicate full range of projections across the ensemble. The baseline frequency of warm days of 10% is represented on the graphs by the dashed line. A full list of CMIP5 models for which data is shown here can be found in supplementary material Table SM21-4.] [INSERT FIGURE 21-8 HERE Figure 21-8: The frequency of 'very wet days' (defined here as the 90th percentile of daily precipitation on wet days during a baseline period of 1961-1990 with wet days defined as days with 1mm of precipitation or more) projected for the 2071-2100 period by 26 CMIP5 GCMs for Asia. Map: Ensemble median frequency of 'very wet days' during 2071-2100 under RCP8.5. Graphs: Box-and-whisker plots indicate the range of regionally-averaged 'very wet day' frequency by 2041-2070 and 2071-2100 under RCPs 4.5 and 8.5 across the 26 CMIP5 models for each SREX sub- regions in Asia Boxes represent inter-quartile range and whiskers indicate full range of projections across the ensemble. The baseline frequency of Very wet days of 10% is represented on the graphs by the dashed line. A full Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 list of CMIP5 models for which data is shown here can be found in supplementary material Table SM21-4. (Note the WMO Expert Team on Climate Change Detection Indices defines very wet days threshold as the 95%-ile daily precipitation event.] _____ END BOX 21-4 HERE _____ 21.3.3.5. Projected Changes in Sea Level Projections of regional sea level changes, both based on the CMIP3 and CMIP5 models, indicate a large regional variability of sea level rise (even more than 100% of the global mean sea level rise) in response to different regional processes (WGI Chapter 13, Section 13.6.5). However, by the end of the 21st century it is very likely that over about 95% of the oceans will undergo sea level rise, with about 70% of coastlines experiencing a sea level rise within 20% of the global value and most regions experiencing sea level fall being located near current and former glaciers and ice sheets (WGI Chapter 13, Section 13.6.5). Some preliminary analysis of the CMIP5 ensembles indicates areas of maximum steric sea level rise in the Northern Atlantic, the northwestern Pacific off the East Asia coasts, the eastern coastal oceanic regions of the Bay of Bengal and the western coastal regions of the Arabian Sea (WGI Chapter 13, Section 13.6.5). 21.3.3.6. Projected Changes in Air Quality Since the AR4 more studies have become available addressing the issue of the effects of both climate and emission changes on air quality. Most of these studies focused on the continental United States and Europe, and utilized both global and regional climate and air quality models run in off-line or coupled mode. Regional modeling studies over the United States or some of its sub-regions include, for example, those of Hogrefe et al. (2004), Knowlton et al. (2004), Steiner et al. (2006), Dawson et al. (2006), Lin et al. (2008), Zhang et al. (2008), Weaver et al. (2009), while examples of global modeling studies include Murazaki and Hess (2006), Stevenson et al. (2006), Shindell et al. (2006), Doherty et al. (2006). Weaver et al. (2009) provide a synthesis of simulated effects of climate change on ozone concentrations in the U.S. using an ensemble of regional and global climate and air quality models, indicating a predominant increase in near-surface ozone concentrations, particularly in the Eastern U.S. (Figure 21-9) mostly tied to higher temperatures and corresponding biogenic emissions. An even greater increase was found in the frequency and intensity of extreme ozone concentration events, which are the most dangerous for human health. Examples of regional studies of air quality changes in response to climate change over Europe include Langner et al. (2005), Forkel and Knocke (2006) , Szopa and Hauglustaine (2007), and Meleux et al. (2007), Kruger et al (2008), Engardt et al. (2009), Carvalho et al. (2010), Andersson and Engardt (2010), Athanassiadou et al. (2010), Katragkou et al. (2010, 2011), Zanis et al. (2011), Huszar et al. (2011), Juda-Rezler et al. (2012). All these studies indicated the potential of large increases in near surface summer ozone concentrations especially in Central and Southern Europe due to much warmer and drier projected summer seasons. [INSERT FIGURE 21-9 HERE Figure 21-9 Mean (top panels) and standard deviation (bottom panels) in future-minus-present (2050s minus 1990s) MDA8 summer ozone concentrations across (left-hand panels) all seven experiments (five regional and 2 global) and for comparison purposes (right hand panels), not including the WSU experiment (which simulated July only conditions). The different experiments use different pollutant emission and SRES GHG emission scenarios. The pollutant emissions are the same in the present and future simulations (from Weaver et al., 2009).] 21.4. Cross-Regional Phenomena Thus far, this chapter has covered climate change-related issues that have a regional expression in one part of the world or another. In principle, these issues can be studied and described, in situ, in the regions in which they occur. However, there is a separate class of issues that transcends regional boundaries and demands a different treatment. In order to understand such cross-regional phenomena, knowledge is required of critical but geographically remote Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 associations and of dynamic cross-boundary flows. The following sections consider some examples of these phenomena, focusing on trade and financial flows and migration. Though these issues are treated in more detail in Part A of this report, they are restated here in Part B to stress the importance of a global perspective in appreciating climate change challenges and potential solutions at the regional scale. 21.4.1. Trade and Financial Flows Global trade and international financial transactions are the motors of modern global economic activity. Their role as key instruments for implementing mitigation and adaptation policies is explored in detail in chapters 14-17 and in the Working Group III report Stavins et al., 2014; Gupta et al., 2014). They are also inextricably linked to climate change (WTO-UNEP, 2009) through a number of other interrelated pathways that are expanded here: (i) as a direct or indirect cause of anthropogenic emissions (e.g., Peters et al., 2011), (ii) as contributory factors for regional vulnerability to the impacts of climate change (e.g. Leichenko and O'Brien, 2008), and (iii) through their sensitivity to climate trends and extreme climate events (e.g., Nelson et al., 2009a; Headey, 2011). 21.4.1.1. International Trade and Emissions The contemporary world is highly dependent on trading relationships between countries in the import and export of raw materials, food and fibre commodities and manufactured goods. Bulk transport of these products, whether by air, sea or over land, is now a significant contributor to emissions of greenhouse gases and aerosols (Stavins et al., 2014). Furthermore, the relocation of manufacturing has transferred net emissions via international trade from developed to developing countries (see Figure 21-10), and most developed countries have increased their consumption-based emissions faster than their domestic (territorial) emissions (Peters et al., 2011). This regional transfer of emissions is commonly referred to in climate policy negotiations as "carbon leakage" (Barker et al., 2007), though only a very small portion of this can be attributed to climate policy ("strong carbon leakage"), a substantial majority being due to the effect of non-climate policies on international trade ("weak carbon leakage" Peters, 2010). A particular example of strong carbon leakage concerns the conversion of land use from the production of food to bioenergy crops. These crops sequester carbon otherwise extracted from the ground as fossil fuels, but in the process displace demand for food production to land in other regions, often inducing land clearance and hence an increase in emissions (Searchinger et al., 2008), though the empirical basis for this latter assertion is disputed (see Kline and Dale, 2008). [INSERT FIGURE 21-10 HERE Figure 21-10: Growth rates from 1990-2008 of international trade, its embodied CO2 emissions and net emissions transfers from Annex B and non-Annex B countries compared to other global macrovariables, all indexed to 1990 (Peters et al., 2011). Annex B and non-Annex B Parties to the UNFCCC are listed in the supplementary material.] 21.4.1.2. Trade and Financial Flows as Factors Influencing Vulnerability The increasingly international nature of trade and financial flows (commonly referred to as globalisation), while offering potential benefits for economic development and competitiveness in developing countries, also presents high exposure to climate-related risks for some of the populations already most vulnerable to climate change (Leichenko and O'Brien, 2008). Examples of these risks, explored further in Chapters 7-9, 12, 13 and 19 of Part A, include: Severe impacts of food price spikes in many developing countries (including food riots and increased incidence of child malnutrition) such as occurred in 2008 following shortfalls in staple cereals, due to a coincidence of regional weather extremes (e.g. drought) in producer countries, the reallocation of food crops by some major exporters for use as biofuels (an outcome of climate policy see previous section) and market speculation (Ziervogel and Ericksen, 2010). Prices subsequently fell back as the world economy went into recession, but spiked again in early 2011 for many of the same reasons (Trostle et al., Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2011), with some commentators predicting a period of rising and volatile prices due to increasing demand and competition from biofuels (Godfray et al., 2010). A growing dependence of the rural poor on supplementary income from seasonal urban employment by family members and/or on international financial remittances from migrant workers (Davies et al., 2009). These workers are commonly the first to lose their jobs in times of economic recession, which automatically decreases the resilience of recipient communities in the event of adverse climate conditions.On the other hand, schemes to provide more effective communication with the diaspora in times of severe weather and other extreme events can provide rapid access to resources to aid recovery and reduce vulnerability (Downing, 2012). Some aspects of international disaster relief, especially the provision of emergency food aid over protracted periods, has been cited as an impediment to enhancing adaptive capacity to cope with climate-related hazards in many developing countries (Schipper and Pelling, 2006). Here, international intervention, while well-intentioned to relieve short-term stress, may actually be counter-productive in regard to the building of long-term resilience. 21.4.1.3. Sensitivity of International Trade to Climate Climate trends and extreme climate events can have significant implications for regional resource exploitation and international trade flows. The clearest example of an anticipated, potentially major impact of climate change concerns the opening of Arctic shipping routes as well as exploitation of mineral resources in the exclusive economic zones (EEZs) of Canada, Greenland/Denmark, Norway, Russia and the USA (Figure 21-11, and see chapter 28, section 28.3.4). For instance, the CCSM4 climate and sea ice model has been used to provide projections under RCP4.5, RCP6.0 and RCP8.5 forcing (see Box 21-1) of future accessibility for shipping to the sea ice hazard zone of the Arctic marine environment defined by the International Maritime Organization (Stephenson et al., 2013 Figure 21-11 (central map). Results suggest that moderately ice-strengthened ships (Polar Class 6), which are estimated under baseline (1980-1999) conditions to be able to access annually about 36 % of the IMO zone, would increase this access to 45-48 % by 2011-2030, 58-69% by 2046-2065 and 68-93% by 2080-2099, with almost complete accessibility projected for summer (90-98% in July-October) by the end of the century (Stephenson et al., 2013). The robustness of those findings was confirmed using seven sea ice models in an analysis of optimal sea routes in peak season (September) for 2050-2069 under RCP 4.5 and RCP8.5 forcing (Smith and Stephenson, 2013). All studies imply increased access to the three major cross Arctic routes: the Northwest Passage, Northern Sea Route (which is part of the Northeast Passage), and Trans-Polar Route (Figure 21-11), which could represent significant distance savings for trans-continental shipping currently using routes via the Panama and Suez Canals (Stephenson et al., 2011). Indeed, in 2009 two ice-hardened cargo vessels the Beluga Fraternity and Beluga Foresight became the first to successfully traverse the Northeast Passage from South Korea to the Netherlands, a reduction of 5,500 km and 10 days compared to their traditional 20,000 km route via the Suez Canal, translating into an estimated saving of some $300,000 per ship, including the cost of standby icebreaker assistance (Smith, 2009; Det Norsk Veritas, 2010). A projection using an earlier version of the CCSM sea ice model under the SRES A1B scenario, but offering similar results (with forcing by mid-century lying just below RCP8.5 chapter 1, Figure 1- 5a), is presented in Figure 21-11 (peripheral maps), which also portrays winter transportation routes on frozen ground. These routes are heavily relied upon for supplying remote communities and for activities such as forestry and, in contrast to the shipping routes, are projected to decline in many regions. [INSERT FIGURE 21-11 HERE Figure 21-11: Central map: Marine exclusive environmental zones (EEZs dashed lines) of Canada, Greenland/Denmark, Norway, Russia, and the USA, and location of the Northwest Passage, Northern Sea Route, Trans-Polar Route, and international high seas within the IMO Guidelines Boundary for Arctic shipping (thick black border). After Stephenson et al. (2013). Peripheral monthly maps: Projected change in accessibility of maritime and land-based transportation by mid-century (2045-2059 relative to 2000-2014) using the Arctic Transport Accessibility Model and CCSM3 climate and sea ice estimates assuming an SRES A1B scenario. Dark blue areas denote new maritime access to Polar Class 6 vessels (light icebreaker); white areas remain inaccessible. Red delimits areas of lost winter road potential for ground vehicles exceeding 2 metric tonnes (Stephenson et al., 2011).] Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 A second illustration of how the risk of adverse climate changes may have contributed to anticipatory adaptive actions affecting countries in other regions of the world and potentially influencing commodity markets, relates to the purchase or renting of large tracts of productive land in parts of Africa, Latin America, Central Asia and Southeast Asia by countries in Europe, Africa, the Gulf and South and East Asia (De Schutter, 2009; Cotula et al., 2011; Zoomers, 2011). While there is clearly a profit motive in many of these purchases (i.e., cheap and fertile land and the opportunity to cultivate high value food or biofuel crops), there is also a concern that domestic agricultural production in some countries will be unable to keep pace with rapid growth in domestic demand and changing dietary preferences, especially in agricultural regions affected by frequent shortfalls due to droughts, floods and cyclones (Cotula et al., 2011), or threatened by sea-level rise (Zoomers, 2011). Land acquisition on such a large scale raises a number of ethical issues relating to local access to food and the appropriate and sustainable management of the land (Deininger and Byerlee, 2012). These issues have led the UN Special Rapporteur on the right to food to recommend a list of eleven principles for ensuring informed participation of local communities, adequate benefit sharing and the respect of human rights (De Schutter, 2009). This issue is elaborated with respect to livelihoods and poverty in chapter 13, section 13.4.3.4, and land dispossession is categorised a key risk in chapter 19, section 19.6.2. Extreme climate phenomena that may be harbingers of similar and more frequent events in a warmer world, already exact devastating consequences in some regions that extend well beyond country boundaries. A recent event that disrupted international trade and commodity flows was the severe 2010/2011 flooding in eastern Australia (Giles, 2011; Queensland Floods Commission of Inquiry, 2012; and see chapter 25, Box 25-8), which combined with damaging cyclones in Queensland and Western Australia curtailed numerous mining operations and damaged transportation networks, leading to declines in both thermal and metallurgical coal exports (by 31% and 19%, respectively, relative to the previous quarter ABARES, 2011) with a sharp rise in their monthly price between November 2010 and January 2011 (Index Mundi, 2012). The severe weather was the primary factor contributing to a fall in Australian GDP of 1.2% during January-March 2011 compared with a rise of 0.7% in the preceding three- month period (Australian Bureau of Statistics, 2011). Other examples of how extreme climate events can affect international trade are reported by Oh and Reuveny (2010) and Handmer et al. (2012). 21.4.2. Human Migration There has been considerable debate in recent years around the postulate that anthropogenic climate change and environmental degradation could lead to mass migration (Perch-Nielsen et al., 2008; Feng et al., 2010; Warner, 2010; Black et al., 2011; Government Office for Science, 2011; Assan and Rosenfeld, 2012). The issue is treated at length in Chapters 9, 12 and 19 of Part A, so only a few aspects are touched on here, to highlight the growing significance of migration in all regions of the world. Four possible pathways through which climate change could affect migration are suggested by Martin (2009): 1) Intensification of natural disasters 2) Increased warming and drought that affects agricultural production and access to clean water 3) Sea-level rise, which makes coastal areas and some island states increasingly uninhabitable 4) Competition over natural resources, leading to conflict and displacement of inhabitants. Abundant historical evidence exists to suggest that changes in climatic conditions have been a contributory factor in migration, including large population displacements in the wake of severe events such as Hurricane Katrina in New Orleans, Louisiana in 2005 (Cutter et al., 2012), Hurricane Mitch in Central America in 1998 and the northern Ethiopian famines of the 1980s (McLeman and Smit, 2006). Other examples are provided in Chapter 12, Table 12-3. However, the evidence is not clear cut (Black, 2001), with counter examples also available of migration being limited due to economic hardship (e.g. during the Sahel drought of the mid-1980s in Mali Findley, 1994). The spatial dimension of climate-related migration is most commonly internal to nations (e.g. from affected regions to safer zones Naik, 2009). In this context it is also worth pointing out that internal migration for other (predominantly economic) reasons may actually expose populations to increased climate risk. For instance, there are large cities in developing countries in low elevation coastal zones that are vulnerable to sea-level rise. Increased Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 migration to these cities could exacerbate the problems with the migrants themselves being especially vulnerable (Nordas and Gleditsch, 2007; UNFPA, 2007). Migration can also be international, though this is less common in response to extreme weather events, and where it does happen it usually occurs along well established routes. For example, emigration following Hurricane Mitch tripled from Honduras and increased from Nicaragua by 40%, mainly to the southern States of the US (already a traditional destination for migrants), and was aided by a relaxation of temporary residency requirements by the United States (Naik, 2009). The causal chains and links between climate change and migration are complex and can be difficult to demonstrate (e.g., Perch-Nielsen et al., 2008; Piguet, 2010; Tänzler et al., 2010; ADB, 2012; Oliver-Smith, 2012; Chapter 12, section 12.4; Chapter 9, section 9.3.3.3.1; Chapter 19, section 19.4.2.1), though useful insights can be gained from studying past abandonment of settlements (McLeman, 2011). Thus projecting future climate-related migration remains a challenging research topic (Feng et al., 2010). There are also psychological, symbolic, cultural and emotional aspects to place attachment, which are well documented from other non-climate causes of forced migration, and are also applicable to cases of managed coastal retreat due to sea-level rise (e.g., Agyeman et al., 2009). Forced migration appears to be an emerging issue requiring more scrutiny by governments in organising development co-operation, and to be factored into international policy making as well as international refugee policies. For example, it has been suggested that the National Adaptation Plans of Action (NAPAs) under the UNFCCC, by ignoring transboundary issues (such as water scarcity), and propounding nationally-orientated adaptation actions (e.g. upstream river management, to the detriment of downstream users in neighbouring countries), could potentially be a trigger for conflict, with its inevitable human consequences. Currently there is no category in the United Nations High Commission for Refugees classification system for environmental refugees, but it is possible that this group of refugees will increase in the future and their needs and rights will need to be taken into consideration (Brown, 2008). The Nansen Initiative, put forward jointly by Norway and Switzerland at a 2011 ministerial meeting, pledges "to cooperate with interested states and relevant actors, including UNHCR, to obtain a better understanding of cross-border movements provoked by new factors such as climate change, identify best practices and develop a consensus on how best to protect and assist those affected ", and may eventually result in a soft law or policy framework (Kolmannskog, 2012). However, migration should not always be regarded as a problem; in those circumstances where it contributes to adaptation (e.g. through remittances) it can be part of the solution (Laczko and Aghazarm, 2009). 21.4.3. Migration of Natural Ecosystems One of the more obvious consequences of climate change is the displacement of biogeographical zones and the natural migration of species (see Chapters 4, 6 and 19). General warming of the climate can be expected to result in migration of ecosystems towards higher latitudes and upward into higher elevations (Chapter 4, section 4.3.2.5) or downward to cooler depths in marine environments (Chapter 6, section 6.3.2.1). Species shifts are already occurring in response to recent climate changes in many parts of the world (Rosenzweig et al., 2008), with average poleward shifts in species range boundaries of 6 km per decade being reported (Parmesan et al., 2011). Study of the estimated shifts of climatic zones alone can provide insights into the types of climatic regimes to anticipate under projected future anthropogenic climate change. By grouping different combinations and levels of climatic variables it is possible not only to track the shifts in the zones in which they occur, but also to identify newly emerging combinations of conditions not found at the present day as well as combinations that may not survive global climate change (known respectively as novel and disappearing climates Williams et al., 2007; and see Chapter 19, section 19.5.1). These analyses can help define what types of climatic niches may be available in the future and where they will be located. Such a spatial analogue approach can delimit those regions that might currently or potentially (in the future) be susceptible to invasion by undesirable aquatic (e.g., EPA, 2008) or terrestrial (e.g., Mainka and Howard, 2010) alien species or alternatively might be candidates for targetting translocation (assisted colonisation) of species endangered in their native habitats (e.g., Brooker et al., 2011; Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Thomas, 2011). However, there are many questions about the viability of such actions, including genetic implications (e.g., Weeks et al., 2011), inadvertent transport of pests or pathogens with the introduced stock (e.g., Brooker et al., 2011) and risk of invasiveness (e.g., Mueller and Hellmann, 2008). The ability of species to migrate with climate change must next be judged, in the first instance, against the rate at which the climatic zones shift over space (e.g. Loarie et al., 2009; Burrows et al., 2011; Diffenbaugh and Field, 2013; Chapter 4, section 4.3.2.5). For projecting potential future species shifts, this is the most straightforward part of the calculation. In contrast, the ecological capacity of species to migrate is a highly complex function of factors, including their ability to: Reproduce, propagate or disperse Compete for resources Adapt to different soils, terrain, water quality and daylength Overcome physical barriers (e.g. mountains, water/land obstacles) Contend with obstacles imposed by human activity (e.g. land use, pollution or dams). Conservation policy under a changing climate is largely a matter of promoting the natural adaptation of ecosystems, if this is even feasible for many species given the rapidity of projected climate change. Studies stress the risks of potential mismatching in responses of co-dependent species to climate change (e.g., Schweiger et al., 2012) as well as the importance of maintaining species diversity as insurance for the provision of basic ecosystem services (e.g., Traill et al., 2010; Isbell et al., 2011). Four priorities have been identified for conservation stakeholders to apply to climate change planning and adaptation (Heller and Zavaleta, 2009): (i) regional institutional coordination for reserve planning and management and to improve landscape connectivity; (ii) a broadening of spatial and temporal perspectives in management activities and practice, and actions to enhance system resilience; (iii) mainstreaming of climate change into all conservation planning and actions; and (iv) holistic treatment of multiple threats and global change drivers, also accounting for human communities and cultures. The regional aspects of conservation planning transcend political boundaries, again arguing for a regional (rather than exclusively national) approach to adaptation policy. This issue is elaborated in Chapters 4 (section 4.4.2) and 19 (section 19.4.2.3). 21.5. Analysis and Reliability of Approaches to Regional Impacts, Adaptation, and Vulnerability Studies Assessing climate vulnerability or options for adapting to climate impacts in human and natural systems requires an understanding of all factors influencing the system and how change may be effected within the system or applied to one or more of the external influencing factors. This will require, in general, a wide range of climate and non- climate information and methods to apply this to enhance the adaptive capacity of the system. There are both areas of commonality across and differences between regions in the information and methods and these are explored in this section. It initially focuses on advances in methods to study vulnerability and adaptive capacity and to assess impacts (studies of practical adaptation and the processes of adaptation decision-making are treated in detail in Chapters 14-17 and so not addressed here). This is followed by assessments of new information on, and thinking related to: (a) baseline and recent trends in factors needed to assess vulnerability and define impacts baselines; and (b) future scenarios used to assess impacts, changes in vulnerability and adaptive capacity; and then assessment of the credibility of the various types of information presented. 21.5.1. Analyses of Vulnerability and Adaptive Capacity Multiple approaches exist for assessing vulnerability and for exploring adaptive capacity (Schipper et al 2010, UNFCCC 2008). The choice of method is influenced by objectives and starting point (see Table 21-3) as well as the type of information available. Qualitative assessments usually draw on different methods and inputs from quantitative assessments. Qualitative information cannot always be translated to quantitative information, or vice versa, yet both approaches can sometimes be used to answer the same questions. Indicators, indices and mapping are the most common ways to aggregate the resulting vulnerability and adaptive capacity information to compare across regions (section 21.5.1.1) or to identify "hotspots" (section 21.5.1.2). . Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 21.5.1.1. Indicators and Indices Several attempts have been made to develop vulnerability indicators and indices (Birkmann 2011; Chen et al. 2011; Barr et al 2010; Cardona, 2007; Luers et al 2003; Lawrence et al., 2003, Villa and McLeod, 2002, Downing et al. 2001, Atkins et al., 2000; Moss et al. 2001). Representation on a map or through an index is a common way to depict global vulnerability information and requires quantification of selected variables in order to measure them against a selected baseline, even though quantification of some qualitative information may not be possible (Hinkel, 2011; Edwards et al, 2007; Luers et al 2003). Vulnerability is differentiated according to factors such as gender, age, livelihood or access to social networks, among many other factors (Cardona et al 2012; Wisner et al 2005), which may not be represented accurately through some indicators. One approach used to create regional comparisons is to use indices, which are composites of several indicators thought to contribute to vulnerability, each normalised and sometimes weighted so they can be combined (Rygel et al 2006, Adger et al 2004). The approach has been critiqued extensively because the weights assigned the indicators depends on expert opinion which can result in different regions appearing more or less vulnerable, as Füssel (2010) found in reviewing global vulnerability maps based on different indices. Vulnerability indices developed to date have failed to reflect the dynamic nature of component indicator variables. This is illustrated by the (in)ability to characterise how the selected indicators contribute to determining vulnerability over time. Importantly, the relative importance of the indicator may change from season-to-season (e.g. access to irrigation water) or may gradually or rapidly become obsolete. Hinkel s (2011) review of literature on vulnerability indicators suggests that vulnerability has been confused as a proxy for unsustainable or insufficient development so that simple measurements are seen as sufficient to tell a story about vulnerability. Hinkel (2011) suggests that the simplification of information to create vulnerability indicators is what limits their utility. Indicator systems have also been developed to improve understanding of adaptive capacity. These are used both to measure adaptive capacity and identify entry points for enhancing it (Adaptation Sub-Committee, 2011; Lioubimtseva and Henebry, 2009; Swanson et al., 2007; Adger and Vincent, 2005; Eriksen and Kelly, 2007). For example, the Global Adaptation Index, developed by the Global Adaptation Alliance (GAIN, n.d.) uses a national approach to assess vulnerability to climate change and other global challenges and compare this with a country s "Readiness to improve resilience" (GAIN, n.d.) to assist public and private sectors to prioritise financial investments in adaptation activities. 21.5.1.2. Hotspots A special case of the use of indicators concerns the identification of hotspots, a term original used in the context of biodiversity, where a biodiversity hotspot is a biologically diverse region typically under threat from human activity, climate change or other drivers (Myers, 1988). The term typically relates to a geographical location, which emerges as a concern when multiple layers of information are compiled to define it. In climate change analysis, hotspots are used to indicate locations that stand out in terms of impacts, vulnerability or adaptive capacity (or all three). Examples of hotspot mapping include how climate change can influence disease risk (de Wet et al., 2001), extinctions of endemic species (Malcolm et al., 2006), and disaster risk (Dilley, 2006). Hotspots analysis is used to serve various purposes, such as setting priorities for policy action, identifying focal regions for further research (de Sherbinin, 2013; Dilley, 2006, Ericksen et al., 2011, see www.climatehotmap.org), or, increasingly, helping distinguish priority locations for funding. Examples of the latter purpose include guiding the allocation of global resources to pre-empt, or combat, disease emergence (Jones et al 2008) or funding for disaster risk management (Arnold et al 2005). Because identifying hotspots raises important methodological issues about the limitation of using indicators to integrate quantitative impacts with qualitative dimensions of vulnerability, their use to compare regions leads to a subjective ranking of locations as having priority for climate change investment. This can be controversial and considered politically-motivated (Klein 2009). Certain locations are considered hotspots because of their regional or global importance. These can be defined by population size and growth rate, contributions to regional or global economies, productive significance (e.g., food Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 production) as well as by disaster frequency and magnitude, and projected climate change impacts. The choice of variables may result in different locations being identified as hotspots (Füssel, 2009). For example, the Consultative Group on International Agricultural Research (CGIAR) Research Program on Climate Change Agriculture and Food Security (CCAFS) mapped hotspots of food insecurity and climate change in the tropics (Ericksen et al, 2011) using stunted growth as a proxy for food security, but other variables could also have been selected. Scale matters in representing hotspots and they will look different on a global scale than on a finer scale (Arnold et al., 2006). The rationale for identifying such hotspots is that they may gradually evolve into locations of conflict or disaster, where a combination of factors lead to the degradation of resources and social fabric. Climate change hotspots have been defined as locations where impacts of climate change are "well pronounced and well documented" (UCS 2011). A climate change hotspot can describe (a) a region for which potential climate change impacts on the environment or different activity sectors can be particularly pronounced or (b) a region whose climate is especially responsive to global change (Giorgi 2006). An example of the former is given by Fraser et al (2013), combining hydrological modelling with quantitatively modelled adaptive capacity (defined as the inverse of sensitivity to drought) to identify vulnerability hotspots for wheat and maize. Examples of the latter are given by Giorgi (2006), Diffenbaugh et al. (2008), Giorgi and Bi (2009), Xu et al. (2009), Diffenbaugh and Scherer (2011) and Diffenbaugh and Giorgi (2012) who used different regional climate change indices, including changes in mean and interannual variability of temperature and precipitation and metrics of seasonal extremes, to identify the Mediterranean Basin, Central America, Central and West Africa, the Northern high latitude regions, the Amazon, the southwestern United States, Southeast Asia and the Tibetan Plateau as prominent hot-spots. 21.5.2. Impacts Analyses In recent years, there has been increased scrutiny of the methods and tools applied in impact assessment, especially quantitative models that are used to project the biophysical and socio-economic impacts of future climate change (see chapter 2, 2.3.2.1), but also encompassing qualitative methods, including studies of indigenous knowledge (Galloway McLean, 2010, and see chapter 12, 12.3.3). In an advance from previous assessments, different types of impact models are now being applied for the first time in many regions of the world. This is largely due to burgeoning international development support for climate change vulnerability and adaptation studies (Fankhauser, 2010). It is also related to a surge of interest in regional economic assessments in the wake of the Stern review (Stern, 2007) as well as to the evolution of climate models into Earth system models that incorporate a more realistic representation of land surface processes (Flato et al., 2014) and their increased application to study hydrological (chapter 3, 3.4.1), ecophysiological (chapter 4, 4.3.3) and cryospheric (Vaughan et al., 2014) impacts. Potential impacts have been simulated for single as well as multiple sectors, at spatial scales ranging from site or household to global, and over a range of temporal scales and time horizons (Table 21-5). A majority of impact studies still follow the conventional approach where future impacts are modelled based on a set of assumptions (scenarios) about future climate and socioeconomic conditions (see 21.2.3, left hand side of Table 21-3),. However, an increasing number are being undertaken that follow a "socio-institutional" approach to adaptation planning (Downing, 2012), right hand side of Table 21-3, which emphasises the importance of adaptive flexibility and climate resilience given the often intractable, "deep" uncertainties implicit in many projections of future change (e.g., Donley et al., 2012; Garrett et al., 2013; Gersonius et al., 2013). Impact modelling studies also commonly treat aspects of adaptation, either explicitly as modelled options or implicitly as built-in autonomous responses (Dickinson, 2007; White et al., 2011). Furthermore, as an anthropogenic signature is attributed to ongoing climate changes in many regions (Bindoff et al., 2014), and with growing evidence that these changes are having impacts on natural and human systems in many more regions than reported in the AR4 (chapter 18, Rosenzweig and Neofotis, 2013), it is now possible in some regions and sectors to test impact models projections against observed impacts of recent climate change (e.g. Araújo et al., 2005; Barnett et al., 2008; Lobell et al., 2011). This is also an essential element in the attribution of observed impacts (Chapter 18, 18.3, 18.4, 18.5). Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Uncertainties in and Reliability of Impacts Analyses Literature on uncertainty in impacts analyses has focused mainly on the uncertainties in impacts that result from the uncertainties in future climate (Mearns et al., 2001; Carter et al., 2007), and this literature continues to grow since AR4, particularly in the realm of agriculture and water resources (e.g., Wetterhall et al., 20011; Ferrise et al., 2011; Ficklin et al., 2012; Littell et al., 2011, Osborne et al., 2013), but also in other areas such as flood risk (Ward et al., 2013). Furthermore, research has advanced to establish which future climate uncertainties are most important to the resultant uncertainties about crop yields (e.g., Lobell and Burke, 2008) and to apply future resource uncertainties to adaptation studies (Howden et al., 2007). Use of multiple global or regional model scenarios is now found in many more studies (e.g. Gosling et al., 2011; Bae et al., 2011; Arnell, 2011; Olsson et al., 2011) and the use of probabilistic quantification of climate uncertainties has produced estimates of probabilities of changes in future resources such as agriculture and water (e.g., Watterson and Whetton, 2011; Tebaldi and Lobell, 2008). Some studies have developed probability distributions of future impacts by combining results from multiple climate projections and, sometimes, different emissions scenarios, making different assumptions about the relative weight to give to each scenario (Brekke et al., 2009). Nobrega et al. (2011) apply 6 different GCMs and 4 different SRES emissions scenarios to study the impacts of climate change on water resources in the Rio Grande Basin in Brazil and found that choice of GCM was the major source of uncertainty in terms of river discharge. With an ever-increasing number of impacts projections appearing in the literature and the unprecedented rate and magnitude of climate change projected for many regions, some authors have begun to question both the robustness of the impacts models being applied (e.g., Heikkinen et al., 2006; Fitzpatrick and Hargrove, 2009; Watkiss, 2011a) as well as the methods used to represent key uncertainties in impacts projections (e.g., Arnell, 2011; Rötter et al., 2011; White et al., 2011). This is being addressed through several prominent international research efforts, Agricultural Model Intercomparison and Improvement Project, involving crop and economic models at different scales (AgMIP Rosenzweig et al., 2013), the Carbon Cycle Model Intercomparison Project (C4MIP Friedlingstein et al., 2006; Sitch et al., 2008; Arora et al., 2013) and the Water Model Intercomparison Project (WaterMIP Haddeland et al., 2011). Modelling groups from these projects are also participating in the Inter- Sectoral Impact Model Intercomparison Project, initially focusing on intercomparing global impact models for agriculture, ecosystems, water resources, health and coasts under RCP- and SSP-based scenarios (see Box 21-1) with regional models being considered in a second phase of work (ISI-MIP Schiermeier, 2012). AgMIP results for 27wheat models run at contrasting sites worldwide indicate that projections of yield to the mid-21st century are more sensitive to crop model differences than to global climate model scenario differences (Asseng et al., 2013; Carter, 2013). WaterMIP s analysis of runoff and evapotranspiration from five global hydrologic and six land surface models indicate substantial differences in the models estimates in these key parameters (Haddelenad et al., 2011). Finally, as in climate modelling, researchers are now applying multiple impact model and perturbed parameter ensemble approaches to future projections (e.g., Araújo and New, 2007; Jiang et al., 2007; Palosuo et al., 2011), usually in combination with ensemble climate projections treated discretely (e.g., New et al., 2007; Graux et al., 2013; Tao and Zhang, 2013) or probabilistically (e.g., Luo et al., 2007; Fronzek et al., 2009, 2011; Brgesen and Olesen, 2011; Ferrise et al., 2011; Wetterhall et al., 2011). These new impact MIPs, and similar initiatives, have the common purpose of mobilising the research community to address some long-recognised but pervasive problems encountered in impact modelling. A sample of recent papers illustrate the variety of issues being highlighted, e.g. forest model typology and comparison (Medlyn et al., 2011), crop pest and disease modelling and evaluation (Sutherst et al., 2011; Garrett et al., 2013), modelling responses to extreme weather events (Lobell et al., 2010; Asseng et al., 2013), field experimentation for model calibration and testing (Long et al., 2006; Craufurd et al., 2013) and data quality considerations for model input and calibration (Lobell, 2013). Greater attention is also being paid to methods of economic evaluation of the costs of impacts and adaptation at scales ranging from global (e.g., UNFCCC, 2007; Nelson et al., 2009b; Parry et al., 2009; Fankhauser, 2010; Füssel, 2010; Patt et al., 2010), through regional (e.g., EEA, 2007; World Bank, 2010; Ciscar et al., 2011; Watkiss, 2011b), to national (SEI, 2009; GCAP, 2011) and local level (e.g., Perrels et al., 2010). Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 21.5.3. Development and Application of Baseline and Scenario Information 21.5.3.1. Baseline Information: Context, Current Status, and Recent Advances This section deals with defining baseline information for assessing climate change impacts, adaptation and vulnerability The baseline refers to a reference state or behaviour of a system, e.g. current biodiversity of an ecosystem, or a reference state of factors (e.g. agricultural activity, climate) which influence that system (see Glossary entry). For example, the UNFCCC defines the pre-industrial baseline climate, prior to atmospheric composition changes from its baseline pre-industrial state, as a reference for measuring global average temperature rises. A baseline may be used to characterise average conditions and/or variability during a reference period, or may allude to a single point in time, such as a reference year. It may provide information on physical factors such as climate, sea level or atmospheric composition, or on a range of non-climate factors, such as technological, land -use or socio-economic conditions. In many cases a baseline needs to capture much of system's variability to enable assessment of its vulnerability or to test whether significant changes have taken place. Thus the information used to establish this baseline must account for the variability of the factors influencing the system. In the case of climate factors often this requires 30 years of data (e.g. Jones et al., 1997) and sometimes substantially more (e.g. Kendon et al., 2008). Also temporal and spatial properties of systems will influence the information required. Many depend on high resolution information, for example urban drainage systems (high spatial scales) or temperature sensitive organisms (sub-daily time-scales). This section assesses methods to derive relevant climatic and non-climatic information and its reliability. 21.5.3.1.1. Climate baselines and their credibility Observed weather data are generally used as climate baselines, e.g. with an impacts model to form a relevant impacts baseline, though downscaled climate model data are now being used as well. For example Bell et al. (2012) use dynamically and statistically downscaled hourly rainfall data with a 1km river flow model to generate realistic high resolution baseline river flows. These were then compared with future river flows derived used corresponding downscaled future climate projections to generate projected impacts representing realistic responses to the imposed climate perturbations. This use of high resolution data was important to ensure that changes in climate variability the system was sensitive to were taken into account (see also Hawkins et al., 2013). Underscoring the importance of including the full spectrum of climate variability when assessing climate impacts, Kay and Jones (2012) showed a greater range of projected changes in UK river flows resulted when using high time-resolution (daily rather than monthly) climate data. Thus to develop the baseline of a climate-sensitive system it is important to have a good description of the baseline climate, thus including information on its variability on timescales of days to decades. This has motivated significant efforts to enhance the quality, length and homogeneity of, and make available, observed climate records (also important for monitoring, detecting and attributing observed climate change Hartmann et al., 2014; Rhein et al., 2014; Vaughan et al., 2014; Masson-Delmotte et al., 2014; Bindoff et al., 2014). This has included generating new datasets such as APHRODITE (a gridded rain-gauge based dataset for Asia, Yatagai, et al., 2012), coordinated analyses of regional climate indices and extremes by CLIVAR s ETCCDI (see e.g. Zhang et al., 2011) and data rescue work typified by the ACRE initiative (Allan et al., 2011) resulting in analysis and digitization of many daily or sub-daily weather records from all over the world. Also, estimates of uncertainty in the observations are either being directly calculated, e.g. for the HadCRUT4 near-surface temperature record (Morice et al., 2012), or can be generated from multiple datasets, e.g. for precipitation using datasets such as GPCC (Rudolf et al., 2011), TRMM (Huffman et al., 2010) and APHRODITE, Yatagi et al.; 2012. Significant progress has also been made in developing improved and new global reanalyses. These use climate models constrained by long time-series of observations from across the globe to reconstruct the temporal evolution of weather patterns during the period of the observations. An important new development has been the use of digitized surface pressure data from ACRE by the 20th Century Reanalysis (20CR) project (Compo et al., 2011) covering 1871 to the present day. 20CR provides the basis for estimating historical climate variability from the sub- daily to the multi-decadal timescale (Figure 21-12) at any location. It can be used directly, or via downscaling, to Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 develop estimates of the baseline sensitivity of a system to climate and addressing related issues such as establishing links between historical climate events and their impacts. Other advances in reanalyses (http://reanalyses.org) have focused on developing higher quality reconstructions for the recent past. They include a new European Centre for Medium Range Weather Forecasts Reanalyses (ERA) dataset, ERA-Interim (Dee et al., 2011) and the NASA Modern Era Reanalysis for Research and Applications (MERRA Rienecker et al., 2011), 1979-present, the NCEP Climate Forecast System Reanalysis (CFSR), 1979-Jan 2010 (Saha et al., 2010) and regional reanalyses such as the North American Regional Reanalysis (NARR) (Mesinger et al., 2006) and EURO4M (http://www.euro4m.eu/). [INSERT FIGURE 21-12 HERE Figure 21-12: Time series of seasonally averaged climate indices representing three modes of large-scale climate variability: (a) the tropical September to January Pacific Walker Circulation (PWC); (b) the December to March North Atlantic Oscillation (NAO); (c) the December to March Pacific North America (PNA) pattern. Indices (as defined in Brönnimann etal. (2009) are calculated (with respect to the overlapping 1989 1999 period) from various observed, reanalysis and model sources: statistical reconstructions of the PWC, the PNA and the NAO, see Brönnimann et al. (2009) for details, (all cyan); 20CR (pink); NCEP NCAR reanalyses (NNR; dark blue); ERA-40 (green); ERA-Interim (orange). The black line and grey shading represent the ensemble mean and spread from a climate model ensemble with a lower boundary condition of observed seas-surface temperatures and sea-ice from the HadISST dataset (Rayner et al. 2003), see Brönnimann etal. (2009) for details. The model results provide a measure of the predictability of these modes of variability from sea-surface temperature and sea-ice alone and demonstrate that the reanalyses have significantly higher skill in reproduces these modes of variability.] In many regions high temporal and spatial resolution baseline climate information is not available (e.g. Washington et al. 2006, World Weather Watch 2005). Recent reanalyses may provide globally complete and temporally detailed reconstructions of the climate of the recent past but generally lack the spatial resolution or have significant biases (Thorne and Vose, 2010; Dee et al., 2011; Cerezo-Mota et al., 2011). Downscaling the reanalyses can be used with available observations to estimate the error in the resulting reconstructions which can often be significant (Duryan et al., 2010; Mearns et al., 2012). Advances in this area are expected through the WCRP-sponsored Coordinated Regional Downscaling Experiment (CORDEX) project (http://wcrp.ipsl.jussieu.fr/SF_RCD_CORDEX.html; Giorgi et al., 2009) which includes downscaling ERA-Interim over all land and enclosed sea areas (e.g. Nikulin et al. 2012). 21.5.3.1.2. Non-climatic baselines and their credibility Climate-sensitive systems can be influenced by many non-climatic factors, so information on the baseline state of these factors is also commonly required (Carter et al., 2001; Carter et al., 2007). Examples of physical non-climatic factors include availability of irrigation systems, effectiveness of disease prevention or flood protection. Examples of socio-economic factors include levels of social, educational and economic development, political/governance background and available technology. Significant work has been undertaken to collect and make this information available. Local and national governments and international agencies (e.g. UN agencies, World Bank) have been collecting data (http://data.worldbank.org/data-catalog) on the human-related factors for many decades and similarly information on technological developments is widely available. Often these factors are evolving quickly and the baseline is taken as the reference state at a particular point in time rather than aggregated over a longer period. In the case of the physical factors, information on many of these have been refined and updated as they are critical inputs to deriving the climate forcings in the Representative Concentration Pathways (RCPs, van Vuuren et al., 2011) used in CMIP5 (Taylor et al., 2012). This includes updated information on land-use change (Hurtt et al., 2011), atmospheric composition (Meinshausen et al., 2011) and aerosols (Grainer et al., 2011, Lamarque et al., 2011). The importance of establishing an appropriate physical baseline is illustrated in a study of potential climate change impacts on flow in the River Thames in the UK over a 126 year period. No long-term trend is seen in annual maximum flows despite increases in temperature and a major change in the seasonal partitioning of rainfall, winter rainfall becoming larger than summer (Marsh 2004). An investigation of the physical environment found that it had been significantly modified as part of river management activities with increases in channel capacity of 30% over 70 years leading to fewer floods. Thus establishing a baseline for river channel capacity explained the current reduced vulnerability of the Thames to flooding. In a study of the potential for crop adaptation (Challinor et al., 2009), the Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 relevant non-climatic factor identified was technological. Detailed field studies demonstrated that the current germplasm included varieties with a wide range of tolerance to higher temperatures (Badigannavar et al., 2002). This established an agricultural technology baseline, current crop properties, which demonstrated the potential to reduce vulnerability in the system to compensate for the projected climate change impact. 21.5.3.2. Development of Projections and Scenarios Since the AR4 there have been several new developments in the realm of scenarios and projections: 1) a new approach to the construction of global scenarios for use in climate change analysis, initiated with the development of representative concentration pathways (RCPs) (see Box 21.1 for a full description); 2) the development and application of a greater number of higher resolution climate scenarios (21.3.3.2); and 3) further use of multiple scenario elements as opposed to use of climate change scenarios only and greater focus on multiple stressors. 21.5.3.2.1. Application of high-resolution future climate information There are now many examples of the generation and application of high resolution climate scenarios for assessing impacts and adaptation planning. These provide information at resolutions relevant for many impacts and adaptation studies but also, particularly with regard to dynamical downscaling, account for higher resolution forcings, such as complex topography (e.g., Salathé et al., 2010) or more detailed land-atmosphere feedbacks such as in West Africa (Taylor et al. 2011). In an analysis of climate impacts including possible adaptations in the Pacific North West of North America (Miles et al., 2010) application of two dynamically downscaled scenarios was particularly useful for the assessment of effects of climate change on storm water infrastructure (Rosenberg et al., 2010). More widely in North America results from NARCCAP have been used to assess impacts of climate change on available wind energy (Pryor and Barthelmie, 2011), road safety (Hambly et al., 2012), hydrology (Burger et al., 2011; Shrestha et al., 2011), forest drought (Williams et al., 2012), and human health (Li et al., 2012). Several European-led projects have generated and applied high resolution climate scenarios to investigate the impacts of climate change over Europe for agriculture, river flooding, human health, and tourism (Christensen et al. 2012) and on energy demand, forest fire risk, wind storms damage, crop yields and water resources (Morse et al., 2009). The UK developed new UK Climate Projections in 2009 (UKCP09) combining the CMIP3, a perturbed physics GCM and a regional climate model ensemble to develop probabilities of changes in temperature and precipitation at a 25 km resolution (Murphy et al., 2009) to determine probabilities of different impacts of climate change and possible adaptations. In general, with all of this work a range of different techniques have been used with little assessment or guidance on the relative merits of each. 21.5.3.2.2. Use of multiple scenario elements and focus on multiple stressors Many more impacts and adaptation studies now use multiple scenario elements, and focus on multiple stressors as opposed to climate change scenarios and effects alone (e.g., 3.3.2, 4.2.4, and 7.1.2). Good examples of use of multiple scenario elements involve studies of climate change and human health considering additional factors such as urban heat island (e.g., Knowlton et al., 2008; Rosenzweig et al., 2009), population increase and expanded urban areas (McCarthy et al., 2010) and population and socio-economic conditions (Watkiss and Hunt, 2012). As these studies are often undertaken at small scales, local scale information on relevant factors may be inconsistent with larger scale scenario elements used in quantifying other stressors. In recognition of this, efforts have been or are being made to downscale the large-scale scenario elements, e.g. the SRES scenarios were downscaled for Europe (van Vuuren and O Neill, 2006) and economic activity information has been downscaled to 0.5 degree grids in some regions (Gaffin et al., 2004; Grübler et al., 2007; van Vuuren et al., 2010) However, this information is far from comprehensive and has not yet been examined carefully in the impacts and vulnerability literature (van Ruijven et al., 2013). Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Typical non-climate stressors include changes in population, migration, land use, economic factors, technological development, social capital, air pollution, and governance structures. They can have independent, synergistic, or antagonistic effects and their importance varies regionally. Land-use and socio-economic changes are stressors of equal importance to climate change for some studies in Latin America (27.2.2.1), numerous changes in addition to climate strongly affect ocean ecosystem health (6.6.1) and in Asia rapid urbanization, industrialization and economic development are identified as major stressors expected to be compounded by climate change in (24.4-7). Most multiple stressor studies are regional or local in scope. For example Ziervogel and Taylor (2008) examined two different villages in South Africa and found that a suite of stressors are present such as high unemployment, health status (e.g., increased concern about AIDs), and access to education with climate change concerns present only in the context of other impacts such as availability of water. In a study on the Great Lakes region, additional stressors included land use change, population increase, and point source pollution (Danz et al., 2007). Mawdesly et al. (2009) considered wildlife management and biodiversity conservation and noted that reducing pressure from other stressors can maximize flexibility for adaptation to climate change. This increased focus on multiple stressors obviously increases the need for a much wider range of data and wider range of projections for the wide range of stressors, across multiple spatial scales. 21.5.3.3. Credibility of Projections and Scenarios 21.5.3.3.1. Credibility of regional climate projections Obtaining robust regional projections of climate change (i.e. at least a clear indication of the direction of change), requires combining projections with detailed analysis and understanding of the drivers of the changes. The most successful example of this is the application of the attribution of observed global and regional temperature changes using global models incorporating known natural and anthropogenic climate forcing factors (Flato et al., 2014, section 10.3). The ability of GCMs to reproduce the observed variations in temperature, the quantification of the influence of the different forcings factors and how well these influences are captured in the models provide confidence that models capture correctly the physical processes driving the changes. This can also provide confidence in projections of precipitation when physically linked to changes in temperature (Rowell and Jones, 2006; Kendon et al. 2010). It is important, especially with precipitation where regional change may appear to differ in direction from one model to another, to distinguish when changes are significant (Tebaldi et al. 2011, Collins et al., 2014b, Box 12.1). Significant future projections of opposite direction are found with neither possibility able to be excluded on the basis of our physical understanding of the drivers of these changes. For example, McSweeney et al. (2012) found that in an ensemble of GCM projections over south-east Asia, all models simulated the important monsoon processes and rainfall well but projected both positive and negative changes in monsoon precipitation and significantly different patterns of change. Model trends or projections may also be inconsistent with trends in available observations and in these cases, their projections are less credible. For example, the magnitude of the significant drying trend seen in the Sahel from the 1960s to the 1990s is not captured by models driven by observed sea-surface temperatures (SSTs) (e.g. Held et al. 2005) despite statistical analysis demonstrating the role of SSTs in driving Sahel rainfall variability. Thus our understanding of the system and its drivers, and their representation in the models, is incomplete, which complicates the interpretation of future projected changes in this region (e.g. Biasutti et al., 2008, Druyan, 2010). It implies that other processes are important and so research is required to identify these and ensure they are correctly represented in the models, without which projections of rainfall changes over this region cannot be considered reliable. 21.5.3.3.1. Credibility regarding socioeconomic scenario elements Cash et al. (2003) distinguish three criteria for linking scientific knowledge to policy action; credibility (scientific adequacy of a policy-relevant study), salience (relevance of a study s findings to the needs of decision-makers), and legitimacy (the perception that the study is respectful of divergent values and beliefs). Studies examining the performance of scenarios in climate change research across all three of these criteria are rare, but a general conclusion has been that much less attention is paid to salience and legitimacy (Hulme and Dessai 2008, Garb et al. Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2008, Girod et al. 2009). Recognising this a new framework for global scenarios has been developed (Box 21-1), providing researchers greater freedom than hitherto for customizing information provided by global scenarios. These innovations may pose challenges for scientific credibility, and it is unclear how difficult it will be to bring independently developed climate and socioeconomic projections together as scenarios in an internally consistent manner, especially when some of these may include fine-scale regional detail (O Neill and Schweizer, 2011; O'Neill et al., 2013). Due to the common practice for scenario development of using narrative descriptions of alternative futures as the inspiration for socioeconomic simulations (the Story and Simulation approach Alcamo 2008) it has been suggested that the exclusion of some details in socioeconomic scenario studies can affect the internal consistency and therefore the overall credibility of a study (e.g. Lloyd and Schweizer, 2013; Schweizer and Kriegler, 2012). Storylines can offer a point of entry for multi-scalar scenario analyses (Rounsevell and Metzger, 2010), and such sub-global scenario studies have been on the rise (Kok et al. 2011, Preston et al. 2011, Sietz et al. 2011, van Ruijven et al., 2013). Environmental scenario exercises crossing geographical scales suggests that linkages between scenarios at different scales can be hard or soft (Zurek and Henrichs 2007), where downscaling (van Vuuren et al. 2010) would be an example of a hard linkage while other similarities between scenarios would be soft linkages. How to apply flexible interpretations of scientific adequacy and maintain scenario credibility is relatively unexplored, and there is thus a need for studies to document best practices in this respect. 21.6. Knowledge Gaps and Research Needs Understanding of the regional nature of climate change, its impacts, regional and cross-regional vulnerabilities, and options for adaptation is still at a rudimentary level. There are both fundamental and methodological research issues in the physical sciences concerned with the projection of regional changes in the climate system and the potential impacts of those changes on various resource sectors and natural systems. Of equal importance, there are also fundamental gaps in our understanding of the determinants of vulnerability and adaptive capacity, thus presenting methodological challenges for projecting how societal vulnerability might evolve as the climate system changes. While development of new scenarios is a part of the underlying research agenda, they will inevitably be limited without further progress in our knowledge of the determinants of vulnerability. Table 21-8 summarizes major research gaps in the physical, ecological, and social sciences that impede the scientific communities progress in understanding the regional context of climate changes, their consequences, and societies responses. [INSERT TABLE 21-8 HERE] Table 21-8: Leading knowledge gaps and related research needs.] Frequently Asked Questions FAQ 21.1: How does this report stand alongside previous assessments for informing regional adaptation? [to be inserted in Section 21.3] The five major Working Group II Assessment Reports produced since 1990 all share a common focus that addresses the environmental and socioeconomic implications of climate change. In a general sense, the earlier assessments are still valid, but the assessments have become much more complete over time, evolving from making very simple, general statements about sectoral impacts, through greater concern with regions regarding observed and projected impacts and associated vulnerabilities, through to an enhanced emphasis on sustainability and equity, with a deeper examination of adaptation options. Finally, in the current report there is a much improved appreciation of the context for regional adaptation and a more explicit treatment of the challenges of decision-making within a risk management framework. Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Obviously one can learn about the latest understanding of regional impacts, vulnerability and adaptation in the context of climate change by looking at the most recent report. This builds on the information presented in previous reports by reporting developments in key topics. New and emergent findings are given prominence, as these may present fresh challenges for decision-makers. Differences with the previous reports are also highlighted whether reinforcing, contradicting or offering new perspectives on earlier findings as these too may have a bearing on past and present decisions. Following its introduction in the Third Assessment Report (TAR), uncertainty language has been available to convey the level of confidence in key conclusions, thus offering an opportunity for calibrated comparison across successive reports. Regional aspects have been addressed in dedicated chapters for major world regions, first defined following the Second Assessment and used with minor variations in the three subsequent assessments. These comprise the continental regions of Africa, Europe, Asia, Australasia, North America, Central and South America, Polar Regions and Small Islands , with a new chapter on The Oceans added for the present assessment. FAQ 21.2: Do local and regional impacts of climate change affect other parts of the world? [to be inserted in Section 21.3.1] Local and regional impacts of climate change, both adverse and beneficial, may indeed have significant ramifications in other parts of the world. Climate change is a global phenomenon, but often expresses itself in local and regional shocks and trends impacting vulnerable systems and communities. These impacts often materialize in the same place as the shock or trend, but also much farther afield, sometimes in completely different parts of the world. Regional interdependencies include both the global physical climate system as well as economic, social and political systems that are becoming increasingly globalised. In the physical climate system, some geophysical impacts can have large-scale repercussions well beyond the regions in which they occur. A well-known example of this is the melting of land-based ice, which is contributing to sea-level rise (and adding to the effects of thermal expansion of the oceans), with implications for low-lying areas far beyond the polar and mountain regions where the melting is taking place. Other local impacts can have wider socio-economic and geopolitical consequences. For instance, extreme weather events in one region may impact production of commodities that are traded internationally, contributing to shortages of supply and hence increased prices to consumers, influencing financial markets and disrupting food security worldwide, with social unrest a possible outcome of food shortages. Another example, in response to longer-term trends is the potential prospect of large-scale migration due to climate change. While hotly contested, this link is already seen in the context of natural disasters, and could become an issue of increasing importance to national and international policy makers. A third example is the shrinkage of Arctic sea ice, opening Arctic shipping routes as well as providing access to valuable mineral resources in the exclusive economic zones of countries bordering the Arctic, with all the associated risks and opportunities. Other examples involving both risks and opportunities include changes of investment flows to regions where future climate change impacts may be beneficial for productivity Finally, some impacts that are entirely local and may have little or no direct effect outside the regions in which they occur still threaten values of global significance, and thus trigger international concern. Examples include humanitarian relief in response to local disasters or conservation of locally threatened and globally valued biodiversity. FAQ 21.3: What regional information should I take into account for climate risk management for the 20 year time horizon? [to be inserted in Section 21.3.2] The fundamental information required for climate risk management is to understand the climate events that put the system being studied at risk and what is the likelihood of these arising. The starting point for assembling this information is a good knowledge of the climate of the recent past including any trends in aspects of these events (e.g. their frequency or intensity). It is also be important to consider that many aspects of the climate are changing, to understand how the future projected changes may influence the characteristics of these events and that these changes will, in general, be regionally variable. However, it should be noted that over the coming 20 years the magnitude of projected changes may not be sufficient to have a large influence the frequency and intensity of these events. Finally, it is also essential to understand which other factors influence the vulnerability of the system. These may be important determinants in managing the risks and also if they are changing at faster rates than the climate then changes in the latter become a secondary issue. Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 For managing climate risks over a 20-year time horizon it is essential to identify the climate variables which the system at risk is vulnerable to. It could be a simple event such as extreme precipitation or a tropical cyclone or a more complex sequence of a late onset of the monsoon coupled with prolonged dry spells within the rainy season. The current vulnerability of the system can then be estimated from historical climate data on these variables including any information on trends in the variables. These historical data would give a good estimate of the vulnerability assuming the record was long enough to provide a large sample of the relevant climate variables and that the reasons for any trends, e.g. clearly resulting from climate change, were understood. It should be noted that in many regions sufficiently long historical records of the relevant climate variables are often not available. It is also important to recognize that many aspect of the climate of the next 20 years will be different from the past. Temperatures are continuing to rise with consequent increases in evaporation and atmospheric humidity and reductions in snow amount and snow season length in many regions. Average precipitation is changing in many regions with both increases and decreases and there is a general tendency for increases in extreme precipitation observed over land areas. There is a general consensus amongst climate projections that further increases in heavy precipitation will be seen as the climate continues to warm and more regions will see significant increases or decreases in average precipitation. In all cases the models project a range of changes for all these variables which are generally different for different regions. Many of these changes may often be relatively small compared to their natural variations but it is the influence of these changes on the specific climate variables which the system is at risk from that is important. Thus information needs to be derived from the projected climate changes on how the characteristics of these variables, e.g. the likelihood of their occurrence or magnitude, will change over the coming 20 years. These projected future characteristics in some cases may be indistinguishable from those historically observed but in other cases some or all models will project significant changes. In the latter situation, the effect of the projected climate changes will then result in a range of changes in either the frequency or magnitude of the climate event, or both. The climate risk management strategy would then need to adapt to accounting for either a greater range or changed magnitude of risk. This implies that in these cases a careful analysis of the implications of projected changes for the specific temporal and spatial characteristic of the climate variables relevant to the system at risk is required. FAQ 21.4: Is the highest resolution climate projection the best to use for performing impacts assessments? [to be inserted in Section 21.5.3.3] A common perception is that higher resolution (i.e., more spatial detail) equates to more useable and robust information. Unfortunately data does not equal information, and more high resolution data does not necessarily translate to more or better information. Hence, while high resolution global climate models (GCMs) and many downscaling methods can provide high resolution data, and add value in, for example, regions of complex topography, it is not a given that there will be more value in the final climate change message. This partially depends upon how the higher resolution data were obtained. For example, simple approaches such as spatial interpolation or adding climate changes from GCMs to observed data fields do increase the spatial resolution but add no new information on high resolution climate change. Nonetheless, these data sets are useful for running impacts models. Many impacts settings are somewhat tuned to a certain resolution, such as the nested size categorizations of hydrologic basins down to watershed size, commonly used in hydrologic modeling. Using dynamical or statistical downscaling methods will add a new high resolution component, providing extra confidence that sub-GCM scale processes are being represented more accurately. However, , there are new errors associated with the additional method applied which need to be considered. More importantly, if downscaling is applied to only one or two GCMs then the resulting high resolution scenarios will not span the full range of projected changes that a large GCM ensemble would indicate are plausible futures. Spanning that full range is important in being able to properly sample the uncertainty of the climate as it applies in an impacts context. Thus for many applications, such as understanding the full envelope of possible impacts resulting from our current best estimates of regional climate change, lower resolution data may be more informative. At the end of the day, no one data set is best, and it is through the integration of multiple sources of information that robust understanding of change is developed. What is important in many climate change impacts contexts is appropriately sampling the full range of known uncertainties, regardless of spatial resolution. It is through the integration of multiple sources of information that robust understanding of change is developed. Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cross-Chapter Box Box CC-RC. Regional Climate Summary Figures [Noah Diffenbaugh (USA), Dáithí Stone (Canada / South Africa / USA), Peter Thorne (USA / Norway / UK), Filippo Giorgi (Italy), Bruce Hewitson (South Africa), Richard Jones (UK), Geert Jan van Oldenborgh (Netherlands)] Information about the likelihood of regional climate change, assessed by WGI, is foundational for the Working Group II assessment of climate-related risks. To help communicate this assessment, the regional chapters of WGII present a coordinated set of regional climate figures, which summarize observed and projected change in annual average temperature and precipitation during the near-term and the longer-term for RCP2.6 and RCP8.5. These WGII regional climate summary figures use the same temperature and precipitation fields that are assessed in WGI Chapter 2 and WGI Chapter 12, with spatial boundaries, uncertainty metrics, and data classes tuned to support the WGII assessment of climate-related risks and options for risk management. Additional details on regional climate and regional climate processes can be found in WGI Chapter 14 and WGI Annex 1. The WGII maps of observed annual temperature and precipitation use the same source data, calculations of data sufficiency, and calculations of trend significance as WGI Chapter 2 and WGI Figures SPM.1 and SPM.2. (A full description of the observational data selection and significance testing can be found in WGI Box 2.2.) Observed trends are determined by linear regression over the 1901-2012 period of MLOST for annual temperature, and over the 1951-2010 period of GPCC for annual precipitation. Data points on the maps are classified into three categories, reflecting the categories used in WGI Figures SPM.1 and SPM.2: 1) Solid colors indicate areas where (i) sufficient data exist to permit a robust estimate of the trend (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period), and (ii) the trend is significant at the 10% level (after accounting for autocorrelation effects on significance testing). 2) Diagonal lines indicate areas where sufficient data exist to permit a robust estimate of the trend, but the trend is not significant at the 10% level. 3) White indicates areas where there are not sufficient data to permit a robust estimate of the trend. The WGII maps of projected annual temperature and precipitation are based on the climate model simulations from Phase 5 of the Coupled Model Intercomparison Project (CMIP5) (Taylor et al., 2012), which also form the basis for the figures presented in WGI (including WGI Chapter 12, Chapter 14, and Annex I). The CMIP5 archive includes output from atmosphere-ocean general circulation models (AOGCMs), AOGCMs with coupled vegetation and/or carbon cycle components, and AOGCMs with coupled atmospheric chemistry components. The number of models from which output is available, and the number of realizations of each model, varies between the different CMIP5 experiments. The WGII regional climate maps use the same source data as WGI Chapter 12 (e.g., Box 12.1 Figure 1), including the WGI multi-model mean values; the WGI individual model values; the WGI measure of baseline ( internal ) variability; and the WGI time periods for the reference (1986-2005), mid-21st-century (2046-2065), and late-21st- century (2081-2100) periods. The full description of the selection of models, the selection of realizations, the definition of internal variability, and the interpolation to a common grid can be found in WGI Chapter 12 and Annex 1. In contrast to Phase 3 of the Coupled Model Intercomparison Project (CMIP3) (Meehl et al., 2007), which used the IPCC SRES emission scenarios (IPCC, 2000), CMIP5 uses the Representative Concentration Pathways (RCPs) (van Vuuren et al., 2011) to characterize possible trajectories of climate forcing over the 21st century. The WGII regional climate projection maps include RCP2.6 and RCP8.5, which represent the high and low end of the RCP range at the end of the 21st century. Projected changes in global mean temperature are similar across the RCPs over the next few decades (Figure RC-1; WGI Fig. 12.5). During this near-term era of committed climate change, risks will evolve as socioeconomic trends interact with the changing climate. In addition, societal responses, particularly adaptations, will influence near-term outcomes. In the second half of the 21st century and beyond, the magnitude of global temperature increase diverges across the RCPs (Figure RC-1; WGI Fig. 12.5). For this longer-term era of climate options, near-term and ongoing mitigation and adaptation, as well as development pathways, will determine the risks Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 of climate change. The benefits of mitigation and adaptation thereby occur over different timeframes, and present- day choices thus affect the risks of climate change throughout the 21st century. [INSERT FIGURE RC-1 HERE Figure RC-1: Observed and projected changes in global annual average temperature. Values are expressed relative to 1986-2005. Black lines show the GISTEMP, NCDC-MLOST, and HadCRUT4.2 estimates from observational measurements. Colored shading denotes the +/-1.64 standard deviation range based on simulations from 32 models for RCP2.6 (blue) and 39 models for RCP8.5 (red). Blue and red lines denote the scenario mean for RCP2.6 and RCP8.5, respectively.] The projection maps plot differences in annual average temperature and precipitation between the future and reference periods (Figure RC-2 and Figure RC-3), categorized into four classes. The classes are constructed based on the IPCC uncertainty guidance, providing a quantitative basis for assigning likelihood (Mastrandrea et al., 2010), with likely defined as 66-100% and very likely defined as 90-100%. The classifications in the WGII regional climate projection figures are based on two aspects of likelihood (e.g., WGI Box 12.1 and Knutti et al. (2010)). The first is the likelihood that projected changes exceed differences arising from internal climate variability (e.g., Tebaldi et al. (2011)). The second is agreement among models on the sign of change (e.g., Christensen et al. (2007) and IPCC (2012)). The four classifications of projected change depicted in the WGII regional climate maps are: 1) Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and >90% of models agree on sign of change. These criteria (and the areas that fall into this category) are identical to the highest-confidence category in WGI Box 12.1. This category supersedes other categories in the WGII regional climate maps. 2) Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability, and >66% of models agree on sign of change. 3) Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. 4) Colors with diagonal lines indicate areas with little or no change, where >66% of models show change less than the baseline variability. It should be noted that areas that fall in this category for the annual average could still exhibit significant change at seasonal, monthly and/or daily timescales. [INSERT FIGURE RC-2 HERE Figure RC-2: Observed and projected changes in annual average temperature. (A) Observed temperature trends from 1901-2012 are determined by linear regression. Trends have been calculated where sufficient data permit a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level (after accounting for autocorrelation effects on significance testing). Diagonal lines indicate areas where change is not significant. Observed data are from WGI AR5 Figures SPM.1 and 2.21. The range of grid-point values is -0.53 to +2.50°C over period. (B) CMIP5 multi-model mean projections of annual average temperature changes for 2046-2065 and 2081-2100 under RCP2.6 and RCP8.5, relative to 1986-2005. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability and >90% of models agree on the sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on the sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on the sign of change. Colors with diagonal lines indicate areas with little or no change, where >66% of models show change less than the baseline variability (although there may be significant change at shorter timescales such as seasons, months, or days). Analysis uses model data from WGI AR5 Figure SPM.8, Box 12.1, and Annex I. The range of grid-point values for the multi-model mean is: +0.19 to +4.08C for mid-21st century of RCP2.6; +0.06 to +3.85C for late-21st century of RCP2.6; +0.70 to +7.04C for mid-21st century of RCP8.5; and +1.38 to +11.71°C for late-21st century of RCP8.5.] Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE RC-3 HERE Figure RC-3: Observed and projected changes in annual average precipitation. (A) Observed precipitation trends from 1951-2010 are determined by linear regression. Trends have been calculated where sufficient data permit a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level (after accounting for autocorrelation effects on significance testing). Diagonal lines indicate areas where change is not significant. Observed data are from WGI AR5 Figures SPM.2. The range of grid-point values is -185 to +111 mm/year/decade. (B) CMIP5 multi-model mean projections of annual average precipitation changes for 2046-2065 and 2081-2100 under RCP2.6 and RCP8.5, relative to 1986-2005. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability and >90% of models agree on the sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on the sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on the sign of change. Colors with diagonal lines indicate areas with little or no change, where >66% of models show change less than the baseline variability (although there may be significant change at shorter timescales such as seasons, months, or days). Analysis uses model data from WGI AR5 Figure SPM.8, Box 12.1, and Annex I. The range of grid-point values for the multi-model mean is: -10 to +24% for mid- 21st century of RCP2.6; -9 to +22% for late-21st century of RCP2.6; -19 to +57% for mid-21st century of RCP8.5; and -34 to +112% for late-21st century of RCP8.5.] Box CC-RC References Christensen, J. H., B. Hewitson, et al. (2007). Regional Climate Projections. Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. S. Solomon, D. Qin, M. Manninget al. Cambridge, United Kingdom and New York, NY, USA, Cambridge University Press. IPCC (2012). Managing the risks of extreme events and disasters to advance climate change adaptation. Cambridge, UK, and New York, NY, USA, Cambridge University Press. IPCC, W. G. I. (2000). Special Report on Emissions Scenarios. Cambridge, UK, Cambridge University Press. Knutti, R., R. Furrer, et al. (2010). Challenges in Combining Projections from Multiple Climate Models. Journal of Climate 23(10): 2739-2758. Mastrandrea, M., C. Field, et al. (2010). Guidance note for lead authors of the IPCC fifth assessment report on consistent treatment of uncertainties. Intergovernmental Panel on Climate Change (IPCC). Meehl, G. A., C. Covey, et al. (2007). The WCRP CMIP3 multimodel dataset - A new era in climate change research. Bulletin of the American Meteorological Society 88(9): 1383-1394. Taylor, K. E., R. J. Stouffer, et al. (2012). An overview of CMIP5 and the experiment design. Bulletin of the American Meteorological Society 93(4): 485-498. Tebaldi, C., J. M. Arblaster, et al. (2011). Mapping model agreement on future climate projections. Geophysical Research Letters 38. van Vuuren, D. P., J. Edmonds, et al. (2011). The representative concentration pathways: an overview. Climatic Change 109(1-2): 5-31. References Adaptation Sub-Committee, 2011: Adapting to Climate Change in the UK. Committee on Climate Change, London, pp. 1-101. ADB, 2005: Climate Proofing: a Risk-based Approach to Adaptation. Pacific Studies Series, Pub. Stock No. 030905, Asian Development Bank, Manila, pp. 191. ADB, 2012: Addressing Climate Change and Migration in Asia and the Pacific. Asian Development Bank, Manila, Philippines, pp. 82. Adger, N.W., 1999: Social vulnerability to climate change and extremes in coastal Vietnam. World Development, 27(2), 249-269. Adger, W.N., N. Brooks, M. Kelly, S. Bentham, and S. Eriksen, 2004: New indicators of Vulnerability and Adaptive Capacity. Tyndall Centre for Climate Change Research, Technical Report 7 (adaptive capacity, country, indicators, Measuring, Vulnerability, vulnerability to climate change). Adger, W.N. and K. Vincent, 2005: Uncertainty in adaptive capacity. C.R. Geoscience, 337, 399. Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Adger, W.N., 2006: Vulnerability. Global Environmental Change, 16(3), 268-281. Agrawala, S., S. Klasen, R. Acosta Moreno, L. Barreto-Gomez, T. Cottier, A.E. Gámez-Vázquez, D. Guan, E.E. Gutierrez-Espeleta, L. Jian, Y.G. Kim, J. Lewis, M. Messouli, M. Rauscher, N. Uddin, and A. Venables, 2014: Regional development and cooperation. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O.,et al. (ed.)]. Agyeman, J., P. Devine-Wright, and J. Prange, 2009: Close to the edge, down by the river? Joining up managed retreat and place attachment in a climate changed world. Environment and Planning A, 41, 509-513. Alcamo, J. (ed.), 2008: Environmental Futures: The Practice of Environmental Scenario Analysis.. Elsevier, San Francisco, CA, pp. 212. Allan, R., P. Brohan, G.P. Compo, R. Stone, J. Luterbacher, and S. Brönnimann, 2011: The International Atmospheric Circulation Reconstructions over the Earth (ACRE) Initiative. Bulletin of the American Meteorological Society, 92(11), 1421-1425. Andersson, C. and M. Engardt, 2010: European ozone in a future climate: importance of changes in dry deposition and isoprene emissions. Journal of Geophysical Research: Atmospheres, 115(D2) (Keywords: climate change;surface ozone;sensitivity). Araújo, M.B., R.J. Whittaker, R.J. Ladle, and M. Erhard, 2005: Reducing uncertainty in projections of extinction risk from climate change. Global Ecology and Biogeography, 14, 529-538. Araújo, M.B. and M. New, 2007: Ensemble forecasting of species distributions. Trends in Ecology and Evolution, 22, 42-47. Arnell, N.W., 2011: Uncertainty in the relationship between climate forcing and hydrological response in UK catchments. Hydrology and Earth System Sciences, 15, 897-912. Arnold, M., R.S. Chen, U. Deichmann, M. Dilley, A.L. Lerner-Lam, R.E. Pullen, and Z. Trohanis (eds.), 2006: Disaster Risk Management Series No. 6. In: Natural Disasters Hotspots: Case Studies. The World Bank, Washington DC, pp. 184. Arnold, M., M. Dilley, U. Deichmann, R.S. Chen, and A.L. Lerner-Lam, 2005: Natural disaster hotspots: a global risk analysis. Disaster Risk Management, 5, 1-145. Arora, V.K., G.J. Boer, P. Friedlingstein, M. Eby, C.D. Jones, J.R. Christian, G. Bonan, L. Bopp, V. Brovkin, P. Cadule, T. Hajima, T. Ilyina, K. Lindsay, J.F. Tjiputra, and T. Wu, 2013: Carbon-concentration and carbon- climate feedbacks in CMIP5 Earth system models. Journal of Climate, 26, 5289-5314. Ashfaq, M., Y. Shi, W. Tung, R.J. Trapp, X. Gao, J.S. Pal, and N.S. Diffenbaugh, 2009: Suppression of south Asian summer monsoon precipitation in the 21st century. Geophysical Research Letters, 36(1). Assan, J.K. and T. Rosenfeld, 2012: Environmentally induced migration, vulnerability and human security: consensus, controversies and conceptual gaps for policy analysis. Journal of International Development, 24, 1046-1057. Asseng, S., F. Ewert, C. Rosenzweig, J.W. Jones, J.L. Hatfield, A.C. Ruane, K.J. Boote, P.J. Thorburn, R.P. Rötter, D. Cammarano, N. Brisson, B. Basso, P. Martre, P.K. Aggarwal1, C. Angulo, P. Bertuzzi, C. Biernath, A.J. Challinor, J. Doltra, S. Gayler, R. Goldberg, R. Grant, L. Heng, J. Hooker, L.A. Hunt, J. Ingwersen, R.C. Izaurralde, K.C. Kersebaum, C. Müller, S. Naresh Kumar, C. Nendel, G. O Leary, J.E. Olesen, T.M. Osborne, T. Palosuo, E. Priesack, D. Ripoche, M.A. Semenov, I. Shcherbak, P. Steduto, C. Stöckle, P. Stratonovitch, T. Streck, I. Supit, F. Tao, M. Travasso, K. Waha, D. Wallach, J.W. White, J.R. Williams, and J. Wolf, 2013: Uncertainty in simulating wheat yields under climate change. Nature Climate Change, 3, 827-832. Athanassiadou, M., J. Baker, D. Carruthers, W. Collins, S. Girnary, D. Hassell, M. Hort, C. Johnson, K. Johnson, R. Jones, D. Thomson, N. Trought, and C. Witham, 2010: An assessment of the impact of climate change on air quality at two UK sites. Atmospheric Environment, 44(16), 1877 (Climate change; Air quality; Urban pollution; Dispersion modelling; Ozone; Nitrogen Oxides; PM10). Atkins, J.P., S. Mazzi, and C.D. Easter, 2000: A commonwealth vulnerability index for developing countries: the position of small states. Economic Paper no. 40. London: Commonwealth Secretariat., London. Badigannavar, A., D. Kale, and G. Murty, 2002: Genetic base and diversity in groundnut genotypes. Plant Breeding, 121(4), 348-353. Bae, D., I. Jung, and D.P. Lettenmaier, 2011: Hydrologic uncertainties in climate change from IPCC AR4 GCM simulations of the Chungju Basin, Korea. Journal of Hydrology, 401(1), 90-105. Barker, T., I. Bashmakov, A. Alharthi, M. Amann, L. Cifuentes, J. Drexhage, M. Duan, O. Edenhofer, B. Flannery, M. Grubb, M. Hoogwijk, F.I. Ibitoye, C.J. Jepma, W.A. Pizer, and K. Yamaji, 2007: Mitigation from a cross- Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 sectoral perspective. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Metz, B., O.R. Davidson, P.R. Bosch, R. Dave, and L.A. Meyer(eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 619-690. Barnett, J. and S. O'Neill, 2010: Maladaptation. Global Environmental Change, 20, 211-213. Barnett, T.P., D.W. Pierce, H.G. Hidalgo, C. Bonfils, B.D. Santer, T. Das, G. Bala, A.W. Wood, T. Nozawa, A.A. Mirin, D.R. Cayan, and M.D. Dettinger, 2008: Human-induced changes in the hydrology of the western United States. Science, 319, 1080-1083. Barr, R., S. Fankhauser, and K. Hamilton, 2010: Adaptation investments: a resource allocation framework. Mitigation and Adaptation Strategies for Global Change, 15(8), 843-858. Bell, V., A. Kay, S. Cole, R. Jones, R. Moore, and N. Reynard, 2012: How might climate change affect river flows across the Thames Basin? An area-wide analysis using the UKCP09 Regional Climate Model ensemble. Journal of Hydrology, 442-443, 89-104. Benestad, R.E., 2011: A new global set of downscaled temperature scenarios. Journal of Climate, 24(8), 2080-2098. Beniston, M., D.B. Stephenson, O.B. Christensen, C.A.T. Ferro, C. Frei, S. Goyette, K. Halsnaes, T. Holt, K. Jylhä, B. Koffi, J. Palutikof, R. Schöll, T. Semmler, and K. Woth, 2007: Future extreme events in European climate: an exploration of regional climate model projections. Climatic Change, 81, 71. Bennett, J., F. Ling, D. Post, M. Grose, S. Corney, B. Graham, G. Holz, J. Katzfey, and N. Bindoff, 2012: High- resolution projections of surface water availability for Tasmania, Australia. Hydrology and Earth System Sciences, 16(5), 1287-1303. Biasutti, M., I. Held, A. Sobel, and A. Giannini, 2008: SST forcings and Sahel rainfall variability in simulations of the twentieth and twenty-first centuries. Journal of Climate, 21(14), 3471-3486. Bindoff, N., P. Stott, K.M. AchutaRao, M. Allen, N. Gillett, D. Gutzler, K. Hansingo, G. Hegerl, Y. Hu, S. Jain, I. Mokhov, J. Overland, J. Perlwitz, R. Sebbari, and X. Zhang, 2014: Detection and attribution of climate change: from global to regional. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Birkmann, J., 2011: Regulation and coupling of society and nature in the context of natural hazards. In: Coping with Global Environmental Change, Disasters and Security. [Brauch, H.B., U. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Shourou et al.(eds.)]. AFES Press, pp. 1103-1127. Black, R., 2001: Environmental refugees: myth or reality?, University of Sussex, pp. 19. Black, R., S.R.G. Bennett, S.M. Thomas, and J.R. Beddington, 2011: Migration as adaptation. Nature, 478, 447-449. Blanco, G., R. Gerlagh, S. Suh, J. Barrett, H. De Conick, C.F.D. Morejon, R. Mathur, N. Nakicenovic, A.O. Ahenkorah, J. Pan, H. Pathak, J. Rice, R. Richels, D. Stern, F.L. Toth, and P. Peter Zhou, 2014: Drivers, trends and mitigation. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O.,et al. (ed.)]. pp. 5. Brgesen, C.D. and J.E. Olesen, 2011: A probabilistic assessment of climate change impacts on yield and nitrogen leaching from winter wheat in Denmark. Natural Hazards and Earth System Science, 11, 2541-2553. Brekke, L.D., E.P. Maurer, J.D. Anderson, M.D. Dettinger, E.S. Townsley, A. Harrison, and T. Pruitt, 2009: Assessing reservoir operations risk under climate change. Water Resources Research, 45(4), W04411. Bronnimann, S., A. Stickler, T. Griesser, A.M. Fischer, A. Grant, T. Ewen, T. Zhou, M. Schraner, E. Rozanov, and T. Peter, 2009: Variability of large-scale atmospheric circulation indices for the Northern Hemisphere during the past 100 years. Meteorologische Zeitschrift, 18(4), 379-396. Brooker, R., A. Britton, A. Gimona, J. Lennon, and N. Littlewood, 2011: Literature review: species translocations as a tool for biodiversity conservation during climate change. In: Scottish Natural Heritage Commissioned Report No.440, pp. 68. Brown, C., A. Greene, P. Block, and A. Giannini, 2008: Review of downscaling methodologies for Africa applications. In: IRI Technical Report 08-05. Technical Report. International Research Institute for Climate and Society., Columbia University, New York, USA. Brown, O., 2008: Migration and Climate Change. International Organization for Migration, Geneva, Switzerland, pp. 60. Bukovsky, M.S., D. Gochis, and L.O. Mearns, 2013: Changes in the North American Monsoon as simulated by the NARCCAP suite of models. Journal of Climate, doi: http://dx.doi.org/10.1175/JCLI-D-12-00538.1. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bürger, G., J. Schulla, and A. Werner, 2011: Estimates of future flow, including extremes, of the Columbia River headwaters. Water Resources Research, 47(10), W10520. Burrows, M.T., D.S. Schoeman, L.B. Buckley, P. Moore, E.S. Poloczanska, K.M. Brander, C. Brown, J.F. Bruno, C.M. Duarte, B.S. Halpern, J. Holding, C.V. Kappel, W. Kiessling, M.I. O'Connor, J.M. Pandolfi, C. Parmesan, F.B. Schwing, W.J. Sydeman, and A.J. Richardson, 2011: The pace of shifting climate in marine and terrestrial ecosystems. Science, 334(6056), 652-655. Cabré, M.F., S.A. Solman, and M.N. Nunez, 2010: Creating regional climate change scenarios over southern South America for the 2020 s and 2050 s using the pattern scaling technique: validity and limitations. Climatic Change, 98, 449. Cameron, C., G. Norrington-Davies, and V. te Velde, Managing Climate Extremes and Disasters in Latin America and the Caribbean: Lessons from the SREX Report. . Campbell, J.D., M.A. Taylor, T.S. Stephenson, R.A. Watson, and F.S. Whyte, 2010: Future climate of the Caribbean from a regional climate model. International Journal of Climatology, 31(12), 1866 (climate change;Caribbean;scenarios;rainfall;temperature). Cardona, O.D., 2007: Indicators of disaster risk and risk management. Inter-American Development Bank Summary Report. Inter-American Development Bank (IADB), Washington, D.C. Cardona, O.D., M.K. van Aalst, J. Birkmann, M. Fordham, G. McGregor, R. Perez, R.S. Pulwarty, E.L.F. Schipper, and B.T. Sinh, 2012: Determinants of risk: exposure and vulnerability In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change (IPCC). [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al.(eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 65- 65-108. Carter, T.R., E.L. La Rovere, R.N. Jones, R. Leemans, L.O. Mearns, N. Nakicenovic, A.B. Pittock, S.M. Semenov, and J. Skea, 2001: Developing and applying scenarios.[McCarthy, J.J., O.F. Canziani, N.A. Leary, D.J. Docken, K.S. White(eds.)]. Proceedings of Climate Change 2001: Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change, Cambridge and New York, pp. 145-190. Carter, T.R., R.N. Jones, X. Lu, S. Bhadwal, C. Conde, L.O. Mearns, B.C. O Neill, M.D.A. Rounsevell, and M.B. Zurek, 2007: New assessment methods and the characterisation of future conditions.[Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, C.E. Hanson(eds.)]. Proceedings of Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, Cambridge, pp. 133 171. Carter, T.R., 2013: Agricultural impacts: Multi-model yield projections. Nature Climate Change, 3, 784-786. Carvalho, A., A. Monteiro, S. Solman, A.I. Miranda, and C. Borrego, 2010: Climate-driven changes in air quality over Europe by the end of the 21st century, with special reference to Portugal. Environmental Science & Policy, 13(6), 445 (Air quality modelling; Regional climate change; Meteorological fields; Ozone; Particulate matter). Cash, D.W., W.C. Clark, F. Alcock, N.M. Dickson, N. Eckley, D.H. Guston, J. Jaeger, and R.B. Mitchell, 2003: Knowledge systems for sustainable development. Proceedings of the National Academy of Sciences, 100(14), 8086-8091. CCRA, 2012: The UK Climate Change Risk Assessment 2012 Evidence Report. [HR Wallingford, AMEC Environment & Infrastructure UK Ltd (formerly Entec UK Ltd), The Met Office, Collingwood Environmental Planning, Alexander Ballard Ltd, Paul Watkiss Associates,Metroeconomica (ed.)]. Defra, London, pp. 464. Cerezo-Mota, R., M. Allen, and R. Jones, 2011: Mechanisms controlling precipitation in the northern portion of the North American monsoon. Journal of Climate, 24(11), 2771-2783. Challinor, A., T. Wheeler, D. Hemming, and H. Upadhyaya, 2009: Crop yield simulations using a perturbed crop and climate parameter ensemble: sensitivity to temperature and potential for genotypic adaptation to climate change. Clim.Res, 38, 117-127. Chen, I.C., J.K. Hill, R. Ohlemuller, D.B. Roy, and C.D. Thomas, 2011: Rapid range shifts of species associated with high levels of climate warming. Science, 333, 1024-1026. Christensen, J.H. and O.B. Christensen, 2007: A summary of the PRUDENCE model projections of changes in European climate by the end of this century. Climatic Change, 81, 7-30. Christensen, O.B., C.M. Goodess, and J. Ciscar, 2012: Methodological framework of the PESETA project on the impacts of climate change in Europe. Climatic Change, 112(1), 7-28. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Christensen, J.H., B. Hewitson, A. Busuioc, A. Chen, X. Gao, I. Held, R. Jones, W.-. Kwon, R. Laprise, V.M. Rueda, L.O. Mearns, C.G. Menéndez, J. Räisänen, A. Rinke, R.K. Kolli, A. Sarr, and P. Whetton, 2007: Regional climate projections.[Solomon, S., D. Qin, M. Manning(eds.)]. Proceedings of Climate Change 2007: The Physical Science Basis. Working Group I Contribution to the Intergovernmental Panel on Climate Change Fourth Assessment Report, pp. 847-940. Church, J.A., P.U. Clark, A. Cazenave, J. Gregory, S. Jevrejeva, A. Levermann, M. Merrifield, G. Milne, R.S. Nerem, P. Nunn, A. Payne, W. Tad Pfeffer, D. Stammer, and A. Unnikrishnan, 2014: Sea level change. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Ciais, P., C. Sabine, G. Bala, L. Bopp, V. Brovkin, J. Canadell, A. Chhabra, R. DeFries, J. Galloway, M. Heimann, C. Jones, C. Le Quéré, R. Myneni, S. Piao, and P. Thornton, 2014: Carbon and other biogeochemical cycles. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Ciscar, J.-., A. Iglesias, L. Feyen, L. Szabó, D. Van Regemorter, B. Amelung, R. Nicholls, P. Watkiss, O.B. Christensen, R. Dankers, L. Garrote, C.M. Goodess, A. Hunt, A. Moreno, J. Richards, and A. Soria, 2011: Physical and economic consequences of climate change in Europe. Proceedings of the National Academy of Sciences, 108(7), 2678-2683. Collins, M., J. Arblaster, J.H. Christensen, J. Marozke, G.J. van Oldenborgh, S. Power, M. Rummukainen, and T. Zhou, 2014: Annex I: Atlas of Global and Regional Climate Projections. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Collins, M., R. Knutti, J. Arblaster, J.-. Dufresne, T. Fichefet, P. Friedlingstein, X. Gao, W. Gutowski, T. Johns, G. Krinner, M. Shongwe, C. Tebaldi, A. Weaver, and M. Wehner, 2014: Long-term climate change: projections, commitments and irreversibility. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Comiso, J.C., D.G. Vaughan, I. Allison, J. Carrasco, G. Kaser, R. Kwok, P. Mote, T. Murray, F. Paul, J. Ren, E. Rignot, O. Solomina, K. Steffen, and T. Zhang, 2014: Observations: Cryosphere. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Compo, G., J. Whitaker, P. Sardeshmukh, N. Matsui, R. Allan, X. Yin, B. Gleason, R. Vose, G. Rutledge, and P. Bessemoulin, 2011: The twentieth century reanalysis project. Quarterly Journal of the Royal Meteorological Society, 137(654), 1-28. Coppola, E. and F. Giorgi, 2010: An assessment of temperature and precipitation change projections over Italy from recent global and regional climate model simulations. . International Journal of Climatology, 30, 11. Cotula, L., S. Vermeulen, P. Mathieu, and C. Toulmin, 2011: Agricultural investment and international land deals: evidence from a multi-country study in Africa. Food Security, 3 (Suppl 1), S99-S113. Craufurd, P.Q., V. Vadez, S.V. Krishna Jagadish, P.V. Vara Prasad, and M. Zaman-Allah, 2013: Crop science experiments designed to inform crop modeling. Agricultural and Forest Meteorology, 170, 8-18. Cutter, S., B. Osman-Elasha, J. Campbell, S.-. Cheong, S. McCormick, R. Pulwarty, S. Supratid, and G. Ziervogel, 2012: Managing the risks from climate extremes at the local level. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change (IPCC). [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al.(eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 291-338. Danz, N.P., G.J. Niemi, R.R. Regal, T. Hollenhorst, L.B. Johnson, J.M. Hanowski, R.P. Axler, J.J. Ciborowski, T. Hrabik, and V.J. Brady, 2007: Integrated measures of anthropogenic stress in the US Great Lakes basin. Environmental Management, 39(5), 631-647. Davies, M., K. Oswald, and T. Mitchell, 2009: Climate Change Adaptation, Disaster Risk Reduction and Social Protection. In: Promoting Pro-Poor Growth: Social Protection. Organisation for Economic Co-operation and Development, Paris, pp. 201-217. Dawson, J.P., P.J. Adams, and S.N. Pandis, 2006: Sensitivity of ozone to summertime climate in the Eastern USA: a modeling case study. Atmospheric Environment, 41, 1494. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 De Elia, R. and H. and Cote, 2010: Climate and climate change sensitivity tp model configuration in the Canadian RCM over North America. Meteorologische Zeitschrift, 19, 325-339. De Schutter, O., 2009: Large-scale land acquisitions and leases: A set of core principles and measures to address the human rights challenge. Report presented to the Human Rights Council by the United Nations Special Rapporteur on the right to food, pp. 18. de Sherbinin, A., 2013: Climate change hotspots mapping: what have we learned? Climatic Change, . de Wet, N., W. Ye, S. Hales, R. Warrick, A. Woodward, and P. Weinstein, 2001: Use of a computer model to identify potential hotspots for dengue fever in New Zealand. The New Zealand Medical Journal, 114(1140), 420-422. Dee, D., S. Uppala, A. Simmons, P. Berrisford, P. Poli, S. Kobayashi, U. Andrae, M. Balmaseda, G. Balsamo, and P. Bauer, 2011: The ERA Interim reanalysis: Configuration and performance of the data assimilation system. Quarterly Journal of the Royal Meteorological Society, 137(656), 553-597. Deininger, K. and D. Byerlee, 2012: The rise of large farms in land abundant countries: do they have a future? World Development, 40(4), 701-714. Déqué, M. and S. Somot, 2010: Weighted frequency distributions express modelling uncertainties in the ENSEMBLES regional climate experiments. Climate Research (Open Access for Articles 4 Years Old and Older), 44(2), 195. Déqué, M., D.P. Rowell, D. Luethi, F. Giorgi, J.H. Christensen, B. Rockel, D. Jacob, E. Kjellstroem, M. de Castro, and B.J.J.M. van den Hurk, 2007: An intercomparison of regional climate simulations for Europe: assessing uncertainties in model projections. Climatic Change, 81(1), 53. Dessai, S. and M. Hulme, 2004: Does climate adaptation policy need probabilities? Climate Policy, 4(2), 107-128. Det Norsk Veritas, 2010: Beluga Shipping masters first commercial transit of the Northeast-Passage. Det Norsk Veritas (DNV), . Diallo, I., M.B. Sylla, F. Giorgi, A.T. Gaye, and M. Camara, 2012: Multimodel GCM-RCM Ensemble-Based Projections of Temperature and Precipitation over West Africa for the Early 21st Century. International Journal of Geophysics, 2012. Dickinson, T., 2007: The Compendium of Adaptation Models for Climate Change: First Edition. Adaptation and Impacts Research Division, Environment Canada, Toronto, Canada, pp. 42. Diffenbaugh, N.S., F. Giorgi, and J.S. Pal, 2008: Climate change hotspots in the United States. Geophysical Research Letters, 35(16) (Climate dynamics, Climate impacts, Global climate models, Regional climate change). Diffenbaugh, N.S. and M. Scherer, 2011: Observational and model evidence of global emergence of permanent, unprecedented heat in the 20th and 21st centuries. Climatic Change, 107(3-4). Diffenbaugh, N.S. and F. Giorgi, 2012: Climate change hotspots in the CMIP5 global climate model ensemble. Climatic Change, 114(3-4), 813. Diffenbaugh, N.S. and C.B. Field, 2013: Changes in ecologically critical terrestrial climate conditions. Science, 341, 486-492. Dilley, M., 2006: Chapter 9: Disaster Risk Hotspots: A Project summary. In: Measuring Vulnerability to Natural Hazards - Towards Disaster Resilient Societies. [Birkmann, J. (ed.)]. United Nations University Press, Tokyo, New York, Paris, pp. 182. Doherty, R.M., D.S. Stevenson, C.E. Johnson, W.J. Collins, and M.G. Sanderson, 2006: Tropospheric ozone and El Nino Southern Oscillation: Influence of atmospheric dynamics, biomass burning emissions, and future climate change. Journal of Geophysical Research: Atmospheres, 111(D19) (Climate variability, ENSO, Global climate models, Troposphere: composition and chemistry, Troposphere: constituent transport and chemistry). Dominguez, F., J. Canon, and J. Valdes, 2010: IPCC AR4 climate simulations for the southwestern U.S.: the importance of future ENSO projections. Climatic Change, 99, 499-514. Donley, E.E., R.J. Naiman, and M.D. Marineau, 2012: Strategic planning for instream flow restoration: a case study of potential climate change impacts in the central Columbia River basin. Global Change Biology, 18, 3071- 3086,. Downing, T.E., R. Butterfield, S. Cohen, S. Huq, R. Moss, A. Rahman, Y. Sokona, and L. Stephen, 2001: Vulnerability Indices, Climate Change Impacts and Adaptation. In: UNEP Division of Policy Development and Law. Technical Report. UNEP Division of Policy Development and Law, Nairobi. Downing, T.E., 2012: Views of the frontiers in climate change adaptation economics. Wiley Interdisciplinary Reviews: Climate Change, 3, 161 170. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Druyan, L.M., 2010: Studies of 21st-century precipitation trends over West Africa. International Journal of Climatology, (West African monsoon; Sahel droughts; climate change). Druyan, L.M., J. Feng, K.H. Cook, Y. Xue, M. Fulakeza, S.M. Hagos, A. Konaré, W. Moufouma-Okia, D.P. Rowell, and E.K. Vizy, 2010: The WAMME regional model intercomparison study. Climate Dynamics, 35(1), 175-192. Ebi, K.L., S. Hallegate, T. Kram, N.W. Arnell, T.R. Carter, J. Edmonds, E. Kriegler, R. Mathur, B.C. O'Neill, K. Riahi, H. Winkler, D.P. van Vuuren, and T. Zwickel, 2013: A new scenario framework for climate change research: background, process, and future directions. Climatic Change, doi: 10.1007/s10584-013-0912-3. Edwards, J., M. Gustafsson, and B. Näslund-Landenmark, 2007: Handbook for Vulnerability Mapping. In: EU Asia Pro Eco project. Technical Report. Swedish Rescue Services Agency, Stockholm. EEA, 2007: Climate change: the cost of inaction and the cost of adaptation. European Environment Agency, Copenhagen, Denmark, pp. 67. Elía, R. and H. Cote, 2010: Climate and climate change sensitivity to model configuration in the Canadian RCM over North America. Meteorologische Zeitschrift, 19(4), 325-339. Engardt, M., R. Bergstrom, and C. Andersson, 2009: Climate and emission changes contributing to changes in near- surface ozone in Europe over the coming decades: Results from model studies. Ambio, 38, 453. Engelbrecht, F.A., J.L. Mcgregor, and C.J. Engelbrecht, 2009: Dynamics of the conformal cubic atmospheric model projected climate change signal over southern Africa. International Journal of Climatology, 29(7), 1013 (anthropogenic forcing;climate change;variable-resolution global model;South Indian convergence zone;Indian Ocean High). EPA, 2008: Effects of climate change for aquatic invasive species and implications for management and research. National Center for Environmental Assessment, United States Environmental Protection Agency, Washington, D.C., pp. 336. Ericksen, P., P. Thornton, A. Notenbaert, L. Cramer, P. Jones, and M. Herrero, 2011: Mapping hotspots of climate change and food insecurity in the global tropics. In: CCAFS Report no. 5 . Technical Report. CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Copenhagen, Denmark. Eriksen, S.H. and P.M. Kelly, 2007: Developing credible vulnerability indicators for climate adaptation policy assessment. Mitigation and Adaptation Strategies for Global Change, 12, 495 (adaptation, climate change, climate policy, coping, vulnerability, vulnerability indicators). Fankhauser, S., 2010: The costs of adaptation. Wiley Interdisciplinary Reviews: Climate Change, 1, 23-30. Feng, S., A.B. Krueger, and M. Oppenheimer, 2010: Linkages among climate change, crop yields and Mexico-US cross-border migration. Proceedings of the National Academy of Sciences, 107(32), 14257-14262. Ferrise, R., M. Moriondo, and M. Bindi, 2011: Probabilistic assessments of climate change impacts on durum wheat in the Mediterranean region. Natural Hazards and Earth System Science, 11, 1293-1302. Ficklin, D.L., I.T. Stewart, and E.P. Maurer, 2012: Projections of 21st Century Sierra Nevada Local Hydrologic Flow Components Using an Ensemble of General Circulation Models. JAWRA Journal of the American Water Resources Association, 48(6), 1104-1125. Field, C.B., V. Barros, T.F. Stocker, D. Qin, D. Dokken, K. Ebi, M. Mastrandrea, K. Mach, G. Plattner, and S. Allen, 2012: Managing the risks of extreme events and disasters to advance climate change adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change Cambridge University Press, Cambridge, UK, and New York, NY, USA, . Findley, S., 1994: Does drought increase migration? A study of migration from rural Mali during the 1983-1985 drought. International Migration Review, 28(3), 539-553. Fischer, E.M., S.I. Seneviratne, P.L. Vidale, D. Lüthi, and C. Schär, 2007: Soil Moisture Atmosphere Interactions during the 2003 European Summer Heat Wave. Journal of Climate, 20, 5081 (Soil moisture, Atmosphere land interactions, Summer/warm season, Precipitation). Fisher, B.S., N. Nakiæenoviæ, K. Alfsen, J. Corfee Morlot, F. de la Chesnaye, J.-. Hourcade, K. Jiang, M. Kainuma, E.L. La Rovere, A. Rana, K. Riahi, R. Richels, D.P. van Vuuren, and R. Warren, 2007: Issues related to mitigation in the long-term context. In: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Metz, B., O.R. Davidson, P.R. Bosch, R. Dave, and L.A. Meyer(eds.)]. pp. 169-250. Fitzpatrick, M.C. and W.W. Hargrove, 2009: The projection of species distribution models and the problem of non- analog climate. Biodiversity and Conservation, 18, 2255-2261. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Flato, G., J. Marotzke, B. Abiodun, P. Braconnot, S.C. Chou, W. Collins, P. Cox, F. Driouech, S. Emori, V. Eyring, C. Forest, P. Gleckler, E. Guilyardi, C. Jakob, V. Kattsov, C. Reason, and M. Rummukainen, 2014: Evaluation of climate models. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Forkel, R. and R. Knoche, 2006: Regional climate change and its impacts on photooxidant concentrations in southern Germany: simulations with a coupled regional climate-chemistry model. Journal of Geophysical Research, 111. Fraser, E.D.G., E. Simelton, M. Termansen, S.N. Gosling, and A. South, 2013: Vulnerability hotspots : Integrating socio-economic and hydrological models to identify where cereal production may decline in the future due to climate change induced drought. Agricultural and Forest Meteorology, 170, 195-205. Friedlingstein, P., P. Cox, R. Betts, L. Bopp, W. von Bloh, V. Brovkin, P. Cadule, S. Doney, M. Eby, I. Fung, G. Bala, J. John, C. Jones, F. Joos, T. Kato, M. Kawamiya, W. Knorr, I.K. Lindsay, H.D. Matthews, T. Raddatz, P. Rayner, C. Reick, E. Roeckner, K.-. Schnitzler, R. Schnur, K. Strassmann, A.J. Weaver, C. Yoshikawa, and N. Zeng, 2006: Climate-Carbon Cycle Feedback Analysis: Results from the C4MIP Model Intercomparison. Journal of Climate, 19, 3337-3353. Fronzek, S., T.R. Carter, J. Räisänen, L. Ruokolainen, and M. Luoto, 2009: Applying probabilistic projections of climate change with impact models: a case study for sub-arctic palsa mires in Fennoscandia. Climatic Change, 99, 515-534. Fronzek, S., T.R. Carter, and M. Luoto, 2011: Evaluating sources of uncertainty in modelling the impact of probabilistic climate change on sub-arctic palsa mires. Natural Hazards and Earth System Sciences, 11, 2981- 2995. Füssel, H.-., 2009: Review and quantitative analysis of indices of climate change exposure, adaptive capacity, sensitivity, and impacts. Background note to the World Development Report 2010. World Bank, Washington, D.C. Füssel, H.-., 2010: Modeling impacts and adaptation in global IAMs. Wiley Interdisciplinary Reviews: Climate Change, 1, 288-303. Füssel, H.-., 2010: How inequitable is the global distribution of responsibility, capability, and vulnerability to climate change: A comprehensive indicator-based assessment. Global Environmental Change, 20, 597-611. Füssel, H., 2007: Vulnerability: A generally applicable conceptual framework for climate change research. Global Environmental Change, 17(2), 155-167. Gaffin, S.R., C. Rosenzweig, X. Xing, and G. Yetman, 2004: Downscaling and geo-spatial gridding of socio- economic projections from the IPCC Special Report on Emissions Scenarios (SRES). Global Environmental Change, 14, 105-123. Galloway McLean, K., 2010: Advance Guard: Climate Change Impacts, Adaptation, Mitigation and Indigenous Peoples A Compendium of Case Studies. United Nations University Traditional Knowledge Initiative, Darwin, Australia, pp. 124. Gao, X.J., Y. Shi, and F. Giorgi, 2011: A high resolution simulation of climate change over China. Science China Earth Sciences, 54(3), 463. Gao, X.J., J.S. Pal, and F. Giorgi, 2006: Projected changes in mean and extreme precipitation over the Mediterranean region from a high resolution double nested RCM simulation. Geophysical Research Letters, 33(3) (Climate variability, Extreme events, Hydroclimatology, Precipitation). Garb, Y., S. Pulver, and S.D. VanDeveer, 2008: Scenarios in society, society in scenarios: toward a social scientific analysis of storyline-driven environmental modeling. Environmental Research Letters, 3(4), 045015. Garrett, K.A., A.D.M. Dobson, J. Kroschel, B. Natarajan, S. Orlandini, H.E.Z. Tonnang, and C. Valdivia, 2013: The effects of climate variability and the color of weather time series on agricultural diseases and pests, and on decisions for their management. Agricultural and Forest Meteorology, 170, 216-227. GCAP, 2011: The Economics of Climate Change in the United Republic of Tanzania. Reports to Development Partners Group and the UK Department for International Development. Global Climate Adaptation Partnership, Oxford, UK, pp. 34. Gerbaux, M., N.M.J. Hall, N. Dessay, and I. Zin, 2009: The sensitivity of Sahelian runoff to climate change. Hydrological Sciences Journal, 54(1), 5. Gersonius, B., R. Ashley, A. Pathirana, and C. Zevenbergen, 2013: Climate change uncertainty: building flexibility into water and flood risk infrastructure. Climatic Change, 116, 411-423. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Giles, B.D., 2011: The Queensland floods of December 2010/early January 2011 and the media. Weather, 66(2), 55. Giorgi, F., 2006: Climate change hot-spots. Geophysical Research Letters, 33(8) (Climate change and variability, Climatology, Global climate models, Regional climate change). Giorgi, F. and E. Coppola, 2007: European Climate-change Oscillation (ECO). Geophysical Research Letters, 34. Giorgi, F., 2008: A Simple Equation for Regional Climate Change and Associated Uncertainty. Journal of Climate, 21, 1589 (Climate change, Regional models). Giorgi, F. and P. Lionello, 2008: Climate change projections for the Mediterranean region. Global and Planetary Change, 63, 90. Giorgi, F. and X. Bi, 2009: The Time of Emergence (TOE) of GHG-forced precipitation change Hot-Spots. . Geophysical Research Letters, 36. Giorgi, F., E.S. Im, E. Coppola, N.S. Diffenbaugh, X.J. Gao, L. Mariotti, and Y. Shi, 2011: Higher hydroclimatic intensity with global warming. Journal of Climate, 24(20) (Global warming, Hydrology, Climate variability). Giorgi, F., C. Jones, and G.R. Asrar, 2009: Addressing climate information needs at the regional level: the CORDEX framework. World Meteorological Organization (WMO) Bulletin, 58(3), 175. Giorgi, F. and R. Francesco, 2000: Evaluating uncertainties in the prediction of regional climate change. Geophysical Research Letters, 27, 1295-1298. Godfray, H.C.J., J.R. Beddington, I.R. Crute, L. Haddad, D. Lawrence, J.F. Muir, J. Pretty, S. Robinson, S.M. Thomas, and C. Toulmin, 2010: Food security: the challenge of feeding 9 billion people. Science, 327, 812-818. Goergen, K., G. Lippmann, J. Beersma, G. Brahmer, H. Buiteveld, M. Carambia, O. De Keizer, P. Krahe, E. Nilson, R. Lammersen, C. Perrin , and D. Volken, 2010: Assessment of Climate Change Impacts on Discharge in the Rhine River Basin: Results of the RheinBlick2050 Project. Report No. I-23. . International Commission for the Hydrology of the Rhine Basin (CHR), Lelystad. Gosling, S.N., R.G. Taylor, N.W. Arnell, and M.C. Todd, 2011: A comparative analysis of projected impacts of climate change on river runoff from global and catchment-scale hydrological models. Hydrology and Earth System Sciences, 15, 279-294. Granier, C., B. Bessagnet, T. Bond, A. D Angiola, Denier van der Gon, Hugo, G.J. Frost, A. Heil, J.W. Kaiser, S. Kinne, and Z. Klimont, 2011: Evolution of anthropogenic and biomass burning emissions of air pollutants at global and regional scales during the 1980 2010 period. Climatic Change, 109(1), 163-190. Graux, A.-., G. Bellocchi, R. Lardy, and J.-. Soussana, 2013: Ensemble modelling of climate change risks and opportunities for managed grasslands in France. Agricultural and Forest Meteorology, 170, 114-131. Grose, M.R., M.J. Pook, P.C. McIntosh, J.S. Risbey, and N.L. Bindoff, 2012: The simulation of cutoff lows in regional climate models: reliability and projected trends. Climatic Dynamics, 39, 445-459. Grose, M.R., S.P. Corney, J.J. Katzfey, J.C. Bennett, G.K. Holz, C.J. White, and N.L. Bindoff, 2012: A regional response in mean westerly circulation and rainfall to projected climate warming over Tasmania, Australia. Climate Dynamics, 40(7-8), 2035-2048. Grubler, A., N. Nakicenovic, K. Riahi, F. Wagner, G. Fischer, I. Keppo, M. Obersteiner, B. O'Neill, S. Rao, and F. Tubiello, 2007: Integrated assessment of uncertainties in greenhouse gas emissions and their mitigation: Introduction and overview. Technological Forecasting and Social Change, 74(7), 873. Grübler, A., B. O Neill, K. Riahi, V. Chirkov, A. Goujon, P. Kolp, I. Prommer, S. Scherbov, and E. Slentoe, 2007: Regional, national, and spatially explicit scenarios of demographic and economic change based on SRES. Technological Forecasting & Social Change, 74, 980-1029. Gupta, S., J. Harnisch, D.C. Barua, L. Chingambo, P. Frankel, R.J.G. Vázquez, L. Gómez Echeverri, E. Haites, Y. Huang, R. Kopp, B. Lefevre, H.O. Machado Filho, and E. Massetti, 2014: Cross-cutting investment and finance issues. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O.,et al. (ed.)]. Haddeland, I., D.B. Clark, W. Franssen, F. Ludwig, F. Voß, N.W. Arnell, N. Bertrand, M. Best, S. Folwell, D. Gerten, S. Gomes, S.N. Gosling, S. Hagemann, N. Hanasaki, R. Harding, J. Heinke, P. Kabat, S. Koirala, T. Oki, J. Polcher, T. Stacke, P. Viterbo, G.P. Weedon, and P. Yeh, 2011: Multimodel Estimate of the Global Terrestrial Water Balance: Setup and First Results. Journal of Hydrometeorology, 12(5), 869-884. Haensler, A., S. Hagemann, and D. Jacob, 2011: Dynamical downscaling of ERA40 reanalysis data over southern Africa: added value in the simulation of the seasonal rainfall characteristics. International Journal of Climatology, 31(15), 2338-2349. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hall, J.W., S. Brown, R.J. Nicholls, N.F. Pidgeon, and R.T. Watson, 2012: Proportionate adaptation. Nature Climate Change, 2, 833-834. Hallegatte, S., 2009: Strategies to adapt to an uncertain climate change. Global Environmental Change, 19, 240-247. Hambly, D., J. Andrey, B. Mills, and C. Fletcher, 2012: Projected implications of climate change for road safety in Greater Vancouver, Canada. Climatic Change, , 1-17. Handmer, J., Y. Honda, Z.W. Kundzewicz, N. Arnell, G. Benito, J. Hatfield, I.F. Mohamed, P. Peduzzi, S. Wu, B. Sherstyukov, K. Takahashi, and Z. Yan, 2012: Changes in impacts of climate extremes: human systems and ecosystems. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change (IPCC). [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al.(eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 231-290. Harris, G.R., D.M.H. Sexton, B.B.B. Booth, M. Collins, and J.M. Murphy, 2013: Probabilistic projections of transient climate change. Climate Dynamics, 40(11-12), 2937-2972. Hartmann, D.L., A.M.G. Klein Tank, M. Rusticucci, L. Alexander, S. Broennimann, Y.A.-. Charabi, F. Dentener, E. Dlugokencky, D. Easterling, A. Kaplan, B. Soden, P. Thorne, M. Wild, and P. Zhai, 2014: Observations: Atmosphere and Surface. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Hawkins, E. and R. Sutton, 2009: The potential to narrow uncertainty in regional climate predictions. Bulletin of the American Meteorological Society, 90(8), 1095-1107. Hawkins, E. and R. Sutton, 2011: The potential to narrow uncertainty in projections of regional precipitation change. Climate Dynamics, 37(1), 407-418. Hawkins, E., T.M. Osborne, C. Ho K., and A.J. Challinor, 2013: Calibration and bias correction of climate projections for crop modelling: An idealised case study over Europe. Agricultural and Forest Meteorology, 170, 19-31. Hayhoe, K., J. van Dorn, T. Croley, N. Schlegal, and D. Wuebbles, 2010: Regional climate change projections for Chicago and the Great Lakes. Journal of Great Lakes Research, 36(2), 7. Headey, D., 2011: Rethinking the global food crisis: The role of trade shocks. Food Policy, 36(2), 136-146. Heikkinen, R.K., M. Luoto, M.B. Araújo, R. Virkkala, W. Thuiller, and M.T. Sykes, 2006: Methods and uncertainties in bioclimatic envelope modelling under climate change. Progress in Physical Geography, 30(6), 751-777. Held, I., T. Delworth, J. Lu, K. Findell, and T. Knutson, 2005: Simulation of Sahel drought in the 20th and 21st centuries. Proceedings of the National Academy of Sciences of the United States of America, 102(50), 17891- 17896. Heller, N.E. and E.S. Zavaleta, 2009: Biodiversity management in the face of climate change: A review of 22 years of recommendations. Biological Conservation, 142, 14-32. Hellmuth, M.E., S.J. Mason, C. Vaughan, M.K. van Aalst, and R. Choularton, 2011: A Better Climate for Disaster Risk Management. International Reserach Institute for Climate and Society, Columbia University, New York, New York, pp. 118. Hewitson, B.C., and R.G. Crane, 2006: Consensus between GCM climate change projections with empirical downscaling: precipitation downscaling over South Africa. International Journal of Climatology, 26, 1315. Hewitt, C.D., 2005: The ENSEMBLES project: Providing ensemble-based predictions of climate changes and their impacts. EGGS Newsletter, 13, 22. Hewitt, C., S. Mason, and D. Walland, 2012: The Global Framework for Climate Services. Nature Climate Change, 2(12), 831-832. Hinkel, J., 2011: "Indicators of vulnerability and adaptive capacity": Towards a clarification of the science policy interface. Global Environmental Change, 21(1), 198-208. Hirschi, M., S.I. Seneviratne, V. Alexandrov, F. Boberg, C. Boroneant, O.B. Christensen, H. Formayer, B. Orlowsky, and P. Stepanek, 2011: Observational evidence for soil-moisture impact on hot extremes in southeastern Europe. Nature Geoscience, 4, 17. HLT, 2011: Climate Knowledge for Action: A Global Framework for Climate Services Empowering the most vulnerable. Report of the High-Level Taskforce for the Global Framework for Climate Services. WMO, Geneva, Switzerland, pp. 240. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hogrefe, C., B. Lynn, K. Civerolo, J.Y. Ku, J. Rosenthal, C. Rosenzweig, R. Goldberg, S. Gaffin, K. Knowlton, and P.L. Kinney, 2004: Simulating changes in regional air pollution over the eastern United States due to changes in global and regional climate and emissions. Journal of Geophysical Research: Atmospheres, 109(D22). Howden, S.M., J. Soussana, F.N. Tubiello, N. Chhetri, M. Dunlop, and H. Meinke, 2007: Adapting agriculture to climate change. Proceedings of the National Academy of Sciences, 104(50), 19691-19696. Huffman, G.J., R.F. Adler, D.T. Bolvin, E.J. Nelkin, F. Hossain, and M. Gebremichael, 2010: The TRMM multi- satellite precipitation analysis (TMPA). Satellite Rainfall Applications for Surface Hydrology, , 3-22. Hulme, M. and S. Dessai, 2008: Predicting, deciding, learning: can one evaluate the success of national climate scenarios? Environmental Research Letters, 3, 045013. Environmental Research Letters, 3, 045013. Hurtt, G.C., L.P. Chini, S. Frolking, R.A. Betts, J. Feddema, G. Fischer, J.P. Fisk, K. Hibbard, R.A. Houghton, A. Janetos, C.D. Jones, G. Kindermann, T. Kinoshita, K. Klein Goldewijk, K. Riahi, E. Shevliakova, S. Smith, E. Stehfest, A. Thomson, P. Thornton, D.P. van Vuuren, and Y.P. Wang, 2011: Harmonization of land-use scenarios for the period 1500 2100: 600 years of global gridded annual land-use transitions, wood harvest, and resulting secondary lands. Climatic Change, 109(1), 117-161. Huszar, P., K. Juda-Rezler, T. Halenka, H. Chervenkov, D. Syrakov, A.B. Krueger, P. Zanis, D. Melas, E. Katragkou, M. Reizer, W. Trapp, and M. Belda, 2011: Potential climate change impacts on ozone and PM levels over Central and Eastern Europe from high resolution simulations. Climate Research, 50, 51-68. Im, E.-., M.H. Kim, and W.T. Kwon, 2007: Projected change in mean and extreme climate over Korea from a double nested regional climate model simulation. Journal of the Meteorological Society of Japan, 85, 717-732. Im, E.-., J.B. Ahn, W.T. Kwon, and F. Giorgi, 2008: Multi decadal scenario simulation over Korea using a one-way double-nested regional climate model system. Part 2: Future climate projection (2021-2050). . Climate Dynamics, 30, 239-254. Im, E.-., W.J. Gutowski, and F. Giorgi, 2008: Consistent changes in 21st century daily precipitation from regional climate simulations for Korea using two convection parameterizations. Geophysical Research Letters, 35(L14706). Im, E.-., I.-. Jung, H. Chang, D.-. Bae, and W.-. Kwon, 2010: Hydroclimatological response to dynamically downscaled climate change simulations for Korean basins. Climatic Change, 100(3-4), 485-508. Im, E.-. and J.B. Ahn, 2011: On the elevation dependency of present day climate and future change over Korea from a high resolution regional climate simulation. Journal of the Meteorological Society of Japan, 89, 89-100. Im, E.-., I.W. Jung, and D.-. Bae, 2011: The temporal and spatial structures of recent and future trends in extreme indices over Korea from a regional climate projection. International Journal of Climatology, 31, 72-86. IPCC, 1990: Climate Change: The IPCC Response Strategies. Report of Working Group III of the Intergovernmental Panel on Climate Change.[Bernthal, F. (ed.)]. Washington, D.C., USA, pp. 270. IPCC, 1990: Climate Change: The IPCC Scientific Assessment. Report Prepared for IPCC by Working Group I, Intergovernmental Panel on Climate Change.[Houghton, J.T., G.T. Jenkins, J.J. Ephraums(eds.)]. Cambridge, pp. 365. IPCC, 1990: Climate Change: The IPCC Impacts Assessment. Report prepared for IPCC by Working Group II, Intergovernmental Panel on Climate Change.[Tegart, W.J.M., G.W. Sheldon, D.C. Griffiths(eds.)]. Canberra, pp. 278. IPCC (ed.), 1992: Climate Change 1992: The Supplementary Report to the IPCC Impacts Assessment, Intergovernmental Panel on Climate Change. Australian Government Publishing Service, pp. 114. IPCC, 1992: Climate Change 1992: The Supplementary Report to the IPCC Impacts Assessment, Intergovernmental Panel on Climate Change.[Tegart, W.J.M. and G.W. Sheldon(eds.)]. Canberra, pp. 114. IPCC, 1992: Climate Change 1992: The Supplementary Report to the IPCC Scientific Assessment, Intergovernmental Panel on Climate Change.[Houghton, J.T., B.A. Callander, S.K. Varney(eds.)]. Cambridge, UK, pp. 200. IPCC, 1994: IPCC Technical Guidelines for Assessing Climate Change Impacts and Adaptations. Part of the IPCC Special Report to the First Session of the Conference of the Parties to the UN Framework Convention on Climate Change, Working Group II, Intergovernmental Panel on Climate Change. Carter, T.R., M.L. Parry, S. Nishioka and H. Harasawa (Eds), University College London, UK and Center for Global Environmental Research, National Institute for Environmental Studies, Tsukuba, Japan, pp. 59. IPCC, 1994: Climate Change 1994: Radiative Forcing of Climate Change and An Evaluation of the IPCC IS92 Emission Scenarios, Reports of Working Groups I and III of the Intergovernmental Panel on Climate Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Change.[Houghton, J.T., L.G.M. Filho, J. Bruce, H. Lee, B.A. Callander, E. Haites et al.(eds.)]. Cambridge, United Kingdom, pp. 339. IPCC, 1996: Climate Change 1995: The Science of Climate Change. Contribution of Working Group I to the Second Assessment Report of the Intergovernmental Panel on Climate Change.[Houghton, J.T., L.G.M. Filho, B.A. Callander, N. Harris, A. Kattenberg, K. Maskell(eds.)]. Cambridge, pp. 572. IPCC, 1996: Climate Change 1995: Impacts, Adaptations, and Mitigation of Climate Change: Scientific-Technical Analyses. Contribution of Working Group II to the Second Assessment Report of the Intergovernmental Panel on Climate Change.[Watson, R.T., M.C. Zinyowera, R.H. Moss(eds.)]. Cambridge, pp. 880. IPCC, 1996: Climate Change 1995: Economic and Social Dimensions of Climate Change. Contribution of Working Group III to the Second Assessment Report of the Intergovernmental Panel on Climate Change.[Bruce, J.P., H. Lee, E.F. Haites(eds.)]. Cambridge, pp. 448. IPCC, 1998: Land Use, Land-Use Change and Forestry: A Special Report of the IPCC, Intergovernmental Panel on Climate Change.[Watson, R.T., I.R. Noble, B. Bolin, N.H. Ravindranath, D.J. Verardo, D.J. Dokken(eds.)]. Cambridge, United Kingdom and New York, NY, USA, pp. 377. IPCC, 1998: The Regional Impacts of Climate Change: An Assessment of Vulnerability. A Special Report of IPCC Working Group II, Intergovernmental Panel on Climate Change.[Watson, R.T., M.C. Zinyowera, R.H. Moss, D.J. Dokken(eds.)]. Cambridge, United Kingdom, pp. 517. IPCC, 1999: Aviation and the Global Atmosphere: A Special Report of Working Groups I and III of the Intergovernmental Panel on Climate Change.[J.E.Penner, D.H.Lister, D.J.Griggs, D.J.Dokken, M.McFarland(eds.)]. Cambridge, United Kingdom, pp. 373. IPCC, 2000: Special Report on Emissions Scenarios: A Special Report of Working Group III of the Intergovernmental Panel on Climate Change.[Nakiæenoviæ, N., J. Alcamo, G. Davis, B. DeVries, J. Fenhann, S. Gaffin et al.(eds.)]. Cambridge, pp. 600. IPCC, 2000: Methodological and Technological Issues in Technology Transfer. A Special Report of IPCC Working Group III, Intergovernmental Panel on Climate Change.[Metz, B., O.R. Davidson, J. Martens, N.M.v.R. Sascha, L.V.W. McGrory(eds.)]. Cambridge, United Kingdom and New York, NY, USA, pp. 466. IPCC, 2001: Climate Change 2001: The Scientific Basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate Change.[Houghton, J.T., Y. Ding, D.J. Griggs, M. Noguer, P.J. van der Linden, X. Dai et al.(eds.)]. Cambridge and New York, pp. 881. IPCC, 2001: Climate Change 2001: Impacts, Adaptation, and Vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change.[McCarthy, J.J., O.F. Canziani, N.A. Leary, D.J. Docken, K.S. White(eds.)]. Cambridge and New York, pp. 1032. IPCC, 2001: Climate Change 2001: Mitigation. Contribution of Working Group III to the Third Assessment Report of the Intergovernmental Panel on Climate Change.[Metz, B., O. Davidson, R. Swart, J. Pan(eds.)]. Cambridge and New York, pp. 752. IPCC, 2005: IPCC Special Report on Carbon Dioxide Capture and Storage. Prepared by Working Group III of the Intergovernmental Panel on Climate Change.[Metz, B., O. Davidson, H.C.d. Coninck, M. Loos, L.A. Meyer(eds.)]. Cambridge, United Kingdom and New York, NY, USA, pp. 442. IPCC, 2007: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change.[M.L., P., O.F. Canziani, J.P. Palutikof, P.J.v.d. Linden, C.E. Hanson(eds.)]. Cambridge, United Kingdom and New York, NY, USA, pp. 976. IPCC, 2007: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change.[Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt et al.(eds.)]. Cambridge, United Kingdom and New York, NY, USA, pp. 996. IPCC, 2007: Climate Change 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change.[Metz, B., O.R. Davidson, P.R. Bosch, R. Dave, L.A. Meyer(eds.)]. Cambridge, United Kingdom and New York, NY, USA, pp. 851. IPCC (ed.), 2012: Renewable Energy Sources and Climate Change Mitigation. Special Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1076. IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 582. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 IPCC, 2014: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.[Edenhofer, O.,et al. (ed.)]. Cambridge, UK and New York, USA, pp. 2. IPCC, 2014: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change.[Stocker, T.,et al. (ed.)]. Cambridge, UK and New York, USA, pp. 12. IPCC/TEAP, 2005: Special Report on Safeguarding the Ozone Layer and the Global Climate System: Issues Related to Hydrofluorocarbons and Perfluorocarbons. prepared by the Intergovernmental Panel on Climate Change (IPCC) Working Groups I and III and the Montreal Protocol Technology and Economic Assessment Panel (TEAP).[Metz, B., L. Kuijpers, S. Solomon, S.O. Andersen, O. Davidson, J. Pons et al.(eds.)]. Cambridge, United Kingdom and New York, NY, USA, pp. 478. Isbell, F., V. Calcagno, A. Hector, J. Connolly, W.S. Harpole, P.B. Reich, M. Scherer-Lorenzen, B. Schmid, D. Tilman, J. van Ruijven, A. Weigelt, B.J. Wilsey, E.S. Zavaleta, and M. Loreau, 2011: High plant diversity is needed to maintain ecosystem services. Nature, 477, 199-202. Ishizaki, Y., H. Shiogama, S. Emori, T. Yokohata, T. Nozawa, K. Takahashi, T. Ogura, M. Yoshimori, and T. Nagashima, 2013: Dependence of precipitation scaling patterns on emission scenarios for representative concentration pathways. Journal of Climate, in press. Ishizaki, Y., T. Yokohata, S. Emori, H. Shiogama, K. Takahashi, T. Nozawa, T. Nakaegawa, N. Hanasaki, T. Ogura, and M. Yoshimori, 2013: Validation of pattern scaling approach for determining the maximum available renewable freshwater resources. Journal of Hydrometeorology, in press. Ishizaki, Y., H. Shiogama, S. Emori, T. Yokohata, T. Nozawa, T. Ogura, M. Abe, M. Yoshimori, and K. Takahashi, 2012: Temperature scaling pattern dependence on representative concentration pathway emission scenarios. A letter. Climatic Change, 112, 535-546. Jackson, J.E., M.G. Yost, C. Karr, C. Fitzpatrick, B.K. Lamb, S.H. Chung, J. Chen, J. Avise, R.A. Rosenblatt, and R.A. Fenske, 2010: Public health impacts of climate change in Washington State: projected mortality risks due to heat events and air pollution. Climatic Change, 102(1), 159-186. Jacob, D., J. Petersen, B. Eggert, A. Alias, O.B. Christensen, L.M. Bouwer, A. Braun, A. Colette, M. Déqué, and G. Georgievski, 2013: EURO-CORDEX: new high-resolution climate change projections for European impact research. Regional Environmental Change, (DOI: 10.1007/s10113-013-0499-2). Jaeger, E.B. and S.I. Seneviratne, 2011: Impact of soil moisture-atmosphere coupling on European climate extremes and trends in a regional climate model. Climate Dynamics, 36, 1919. Janssen, M.A. and E. Ostrom, 2006: Resilience, vulnerability, and adaptation: A cross-cutting theme of the International Human Dimensions Programme on Global Environmental Change. Global Environmental Change, 16(3), 237-239. Jiang, T., Y.D. Chen, C. Xu, X. Chen, X. Chen, and V.P. Singh, 2007: Comparison of hydrological impacts of climate change simulated by six hydrological models in the Dongjiang Basin, South China. Journal of Hydrology, 336, 316-333. Jiang, T., Y.D. Chen, C. Xu, X. Chen, X. Chen, and V.P. Singh, 2007: Comparison of hydrological impacts of climate change simulated by six hydrological models in the Dongjiang Basin, South China. Journal of Hydrology, 336, 316-333. Jiang, T., Y.D. Chen, C. Xu, X. Chen, X. Chen, and V.P. Singh, 2007: Comparison of hydrological impacts of climate change simulated by six hydrological models in the Dongjiang Basin, South China. Journal of Hydrology, 336, 316-333. Jiang, T., Y.D. Chen, C. Xu, X. Chen, X. Chen, and V.P. Singh, 2007: Comparison of hydrological impacts of climate change simulated by six hydrological models in the Dongjiang Basin, South China. Journal of Hydrology, 336, 316-333. Jones, K.E., N.G. Patel, M.A. Levy, A. Storeygard, D. Balk, J.L. Gittleman, and P. Daszak, 2008: Global trends in emerging infectious diseases. Nature, 45, doi:10.1038/nature06536. Jones, R., J. Murphy, M. Noguer, and A. Keen, 1997: Simulation of climate change over europe using a nested regional climate model. II: Comparison of driving and regional model responses to a doubling of carbon dioxide. Quarterly Journal of the Royal Meteorological Society, 123(538), 265-292. Juda-Rezler, K., M. Reizer, P. Huszar, B.C. Krüger, P. Zanis, D. Syrakov, E. Katragkou, W. Trapp, D. Melas, H. Chervenkov, I. Tegoulias, and T. Halenka, 2012: Modelling the effects of climate change on air quality over Central and Eastern Europe: concept, evaluation and projections. Climate Research, 53(3), 179-203. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kanada, S., M. Nakano, and T. Kato, 2012: Projections of Future Changes in Precipitation and the Vertical Structure of the Frontal Zone during the Baiu Season in the Vicinity of Japan Using a 5-km-mesh Regional Climate Model. Journal of the Meteorological Society of Japan. Ser. II, 90A, 65-86. Kandlikar, M., H. Zerriffi, and C. Ho Lem, 2011: Science, decision-making and development: managing the risks of climate variation in less-industrialized countries. Wiley Interdisciplinary Reviews: Climate Change, 2, 201-219. Karl, T.R., J.M. Melillo, and T.C. Peterson (eds.), 2009: Global Climate Change Impacts in the United States. Cambridge University Press, New York, NY, USA, pp. 189. Katragkou, E., P. Zanis, I. Tegoulias, D. Melas, I. Kioutsioukis, B.C. Krüger, P. Huszar, T. Halenka, and S. Rauscher, 2010: Decadal regional air quality simulations over Europe in present climate: near surface ozone sensitivity to external meteorological forcing. . Atmospheric Chemistry and Physics, 10(23), 11805-11821. Katragkou, E., P. Zanis, I. Kioutsioukis, I. Tegoulias, D. Melas, B.C. Krüger, and E. Coppola, 2011: Future climate change impacts on summer surface ozone from regional climate air quality simulations over Europe. Journal of Geophysical Research: Atmospheres, 116(D22307). Kay, A. and R. Jones, 2012: Comparison of the use of alternative UKCP09 products for modelling the impacts of climate change on flood frequency. Climatic Change, , 1-20. Kendon, E.J., D.P. Rowell, and R.G. Jones, 2010: Mechanisms and reliability of future projected changes in daily precipitation. Climate Dynamics, 35(2), 489-509. Kendon, E.J., D.P. Rowell, R.G. Jones, and E. Buonomo, 2008: Robustness of Future Changes in Local Precipitation Extremes. Journal of Climate, 21(17), 4280-4297. Kienberger, S., T. Blaschke, and R.Z. Zaidi, 2012: A framework for spatio-temporal scales and concepts from different disciplines: the 'vulnerability cube'. Natural Hazards, 68(3), 1343-1369. Kirtman, B., S. Power, A.J. Adedoyin, G. Boer, R. Bojariu, I. Camilloni, F. Doblas-Reyes, A. Fiore, M. Kimoto, G.A. Meehl, M. Prather, A. Sarr, C. Schär, R. Sutton, G.J. van Oldenborgh, G. Vecchi, and H.-. Wang, 2014: Near-term climate change: projections and predictability. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Kitoh, A., T. Ose, K. Kurihara, S. Kusunoki, M. Sugi, and KAKUSHIN Team-3 Modeling Group, 2009: Projection of changes in future weather extremes using super-high-resolution global and regional atmospheric models in the KAKUSHIN Program: Results of preliminary experiments. Hydrological Research Letters, 3, 49-53. Kitoh, A., S. Kusunoki, and T. Nakaegawa, 2011: Climate change projections over South America in the late 21st century with the 20 and 60 km mesh Meteorological Research Institute atmospheric general circulation model (MRI-AGCM). Journal of Geophysical Research, 116. Kjellstrom, E. and K. Ruosteenoja, 2007: Present day and future precipitation in the Baltic Sea region as simulated in a suite of regional climate models. . Climatic Change, 81, 281. Klein, R.J.T., 2009: Identifying countries that are particularly vulnerable to the adverse effects of climate change: an academic or a political challenge? Carbon & Climate Law Review, 3, 284. Klein, R.J.T., R.J. Nicholls, and F. Thomalla, 2003: Resilience to natural hazards: How useful is this concept? Environmental Hazards, 5(1), 35-45. Kline, K.L. and V.H. Dale, 2008: Biofuels: effect on land and fire. Science, 321, 199. Knowlton, K., J.E. Rosenthal, C. Hogrefe, B. Lynn, S. Gaffin, R. Goldberg, C. Rosenzweig, K. Civerolo, J.-. Ku, and P.L. Kinney, 2004: Assessing Ozone-Related Health Impacts under a Changing Climate. Environmental Health Perspectives, 112, 1557. Knowlton, K., C. Hogrefe, B. Lynn, C. Rosenzweig, J. Rosenthal, and P.L. Kinney, 2008: Impacts of heat and ozone on mortality risk in the New York City metropolitan region under a changing climate. Seasonal Forecasts, Climatic Change and Human Health, , 143-160. Knutson, T.R., J.L. McBride, J. Chan, K. Emanual, G. Holland, C. Landsea, I. Held, J.P. Kossin, A.K. Srivastava, and M. Sugi, 2010: Tropical cyclones and climate change. Nature Geoscience, 3, 157. Knutti, R., R. Furrer, C. Tebaldi, J. Cermak, and G. Meehl, 2010: Challenges in combining projections from multiple climate models. Journal of Climate, 23(10), 2739-2758. Knutti, R., R. Furrer, C. Tebaldi, J. Cernak, and G.A. Meehl, 2010: Challenges in combining projections from multiple climate models. Journal of Climate, 23, 2739-2758. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kok, K., M. van Vliet, I. Bärlund, A. Dubel, and J. Sendzimir, 2011: Combining participative backcasting andexploratory scenario development: Experiences from the SCENES project. , 78, 835-851. Technological Forecasting and Social Change, 78, 835-851. Kolmannskog, V., 2012: Climate change, environmental displacement and international law. Journal of International Development, 24, 1071-1081. Kriegler, E., B.C. O Neill, S. Hallegatte, T. Kram, R.J. Lempert, R.H. Moss, and T. Wilbanks, 2012: The need for and use of socio-economic scenarios for climate change analysis: A new approach based on shared socio- economic pathways. Global Environmental Change, 22(4), 807-822. Kriegler, E., J. Edmonds, S. Hallegatte, K.L. Ebi, T. Kram, K. Riahi, H. Winkler, and D. van Vuuren, 2013: A new scenario framework for climate change research: The concept of Shared Policy Assumptions. Climatic Change, In press. Krüger, B.C., E. Katragkou, I. Tegoulias, P. Zanis, D. Melas, E. Coppola, S. Rauscher, P. Huszar, and T. Halenka, 2008: Regional photochemical model calculations for Europe concerning ozone levels in a changing climate. Idojaras, 112(3-4), 285. Kumar, P., A. Wiltshire, C. Mathison, S. Asharaf, B. Ahrens, P. Lucas-Picher, J.H. Christensen, A. Gobiet, F. Saeed, and S. Hagemann, 2013 (in press): Downscaled climate change projections with uncertainty assessment over India using a high resolution multi-model approach. Science of the Total Environment, . Kunreuther, H., S. Gupta, V. Bosetti, R. Cooke, M.H. Duong, H. Held, J. Llanes-Regueiro, A. Patt, E. Shittu, and E. Weber, 2014: Integrated risk and uncertainty assessment of climate change response policies. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O.,et al. (ed.)]. pp. 2. Laczko, F. and C. Aghazarm (eds.), 2009: Migration, Environment and Climate Change: Assessing the Evidence. International Organization for Migration, Geneva, Switzerland, pp. 441. Lamarque, J., G.P. Kyle, M. Meinshausen, K. Riahi, S.J. Smith, D.P. van Vuuren, A.J. Conley, and F. Vitt, 2011: Global and regional evolution of short-lived radiatively-active gases and aerosols in the Representative Concentration Pathways. Climatic Change, 109(1), 191-212. Langner, J.R., R. Bergstrom, and V. Foltescu, 2005: Impact of climate change on surface ozone and deposition of sulphur and nitrogen in Europe. Atmospheric Environment, 39, 1129. Lassen, M.K., K.D. Nielsen, K. Richardson, K. Garde, and L. Schlüter, 2010: The effects of temperature increases on a temperate phytoplankton community A mesocosm climate change scenario. Journal of Experimental Marine Biology and Ecology, 383, 79-88. Lawrence, P., J.R. Meigh, and C.A. Sullivan, 2003: The Water Poverty Index: An International Comparison. . In: Keele Economics Department Working Paper. Technical. Keele Economics Department Working Paper, Staffordshire, UK. Leichenko, R. and K. O Brien, 2008: Environmental Change and Globalization: Double Exposures. Oxford University Press, New York, NY, USA, pp. 167. Lenderink, G., A. van Ulden, B. van den Hurk, and E. van Meijgaard, 2007: Summertime interannual temperature variability in an ensemble of regional model simulations: analysis of the surface energy budget. Climatic Change, 81, 233. Li, B., S. Sain, L.O. Mearns, H.A. Anderson, S. Kovats, K.L. Ebi, M.Y. Bekkedal, M.S. Kanarek, and J.A. Patz, 2012: The impact of extreme heat on morbidity in Milwaukee, Wisconsin. Climatic Change, 110(3), 959-976. Lin, J,-T, K. O. Patten, K. Hayhoe, X.-. Liang, and D.J. Wuebbles, 2008: Effects of Future Climate and Biogenic Emissions Changes on Surface Ozone over the United States and China. Journal of Applied Meteorology and Climatology, 47, 1888. Lioubimtseva, E. and G.M. Henebry, 2009: Climate and environmental change in arid Central Asia: Impacts, vulnerability, and adaptations. Journal of Arid Environments, 73(11), 963. Littell, J.S., D. McKenzie, B.K. Kerns, S. Cushman, and C.G. Shaw, 2011: Managing uncertainty in climate-driven ecological models to inform adaptation to climate change. Ecosphere, 2(9), art102. Liu, C., D. Golding, and G. Gong, 2008: Farmers coping response to the low flows in the lower Yellow River: A case study of temporal dimensions of vulnerability. Global Environmental Change, 8, 543-553. Lloyd, E.A. and V.J. Schweizer, 2013: Objectivity and a comparison of methodological scenario approaches for climate change research. Synthese, DOI: Loarie, S.R., P.B. Duffy, H. Hamilton, G.P. Asner, C.B. Field, and D.D. Ackerly, 2009: The velocity of climate change. Nature, 462, 1052-1055. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Lobell, D.B. and M.B. Burke, 2008: Why are agricultural impacts of climate change so uncertain? The importance of temperature relative to precipitation. Environmental Research Letters, 3(3), 034007. Lobell, D.B., M. Bänziger, C. Magorokosho, and B. Vivek, 2010: Nonlinear heat effects on African maize as evidenced by historical yield trials. Nature Climate Change, 1, 42-45. Lobell, D.B., W. Schlenker, and J. Costa-Roberts, 2011: Climate trends and global crop production since 1980. Science, 333, 616-620. Lobell, D.B., 2013: Errors in climate datasets and their effects on statistical crop models. Agricultural and Forest Meteorology, 170, 58-66. Long, S.P., E.A. Ainsworth, A.D.B. Leakey, J. Nösberger, and D.R. Ort, 2006: Food for thought: lower-than- expected crop yield stimulation with rising CO2 concentrations. Science, 312, 1918-1921. Luers, A.L., D.B. Lobell, L. Sklar, C. Addams, and P.A. Matson, 2003: A method for quantifying vulnerability, applied to the Yaqui Valley, Mexico. Global Environmental Change, 13, 255. Lumsden, T.G., R.E. Schulze, and B.C. Hewitson, 2009: Evaluation of potential changes in hydrologically relevant statistics of rainfall in Southern Africa under conditions of climate change. Water SA, 35(5), 649-656. Luo, Q., W. Bellotti, M. Williams, I. Cooper, and B. Bryan, 2007: Risk analysis of possible impacts of climate change on South Australian wheat production. Climatic Change, 85, 89-101. Mainka, S.A. and G.W. Howard, 2010: Climate change and invasive species: double jeopardy. Integrative Zoology, 5, 102-111. Malcolm, J.R., L. Canran, R.P. Neilson, L. Hansen, and L. Hannah, 2006: Global warming and extinctions of endemic species from biodiversity hotspots. Conservation Biology, 20(2), 538-548. Malone, E.L. and N.L. Engle, 2011: Evaluating regional vulnerability to climate change: purposes and methods. Wiley Interdisciplinary Reviews: Climate Change, 2, 462-474. Maraun, D., F. Wetterhall, A. Ireson, R. Chandler, E. Kendon, M. Widmann, S. Brienen, H. Rust, T. Sauter, and M. Themeßl, 2010: Precipitation downscaling under climate change: Recent developments to bridge the gap between dynamical models and the end user. Reviews of Geophysics, 48(3), RG3003. Marengo, J.A., T. Ambrizzi, R.P. da Rocha, L.M. Alves, S.V. Cuadra, M.C. Valverde, R.R. Torres, D.C. Santos, and S.E.T. Ferraz, 2010: Future change of climate in South America in the late twenty- rst century: intercomparison of scenarios from three regional climate models. Climate Dynamics, 35, 1089. Marengo, J.A., R. Jones, L.M. Alves, and M.C. Valverde, 2009: Future change of temperature and precipitation extremes in South America as derived from the PRECIS regional climate modeling system. International Journal of Climatology, 29(15), 2241-2255. Mariotti, L., E. Coppola, M.B. Sylla, F. Giorgi, and C. Piani, 2011: Regional climate model simulation of projected 21st century climate change over an all-Africa domain: Comparison analysis of nested and driving model results. Journal of Geophysical Reseaarch, 116(D15). Marsh, T.J., 2004: The January 2003 flood on the Thames. Weather, 59(3), 59-62. Martin, S.F., 2009: Managing environmentally induced migration.[Laczko, F. and C. Aghazarm(eds.)]. Proceedings of Migration, Environment and Climate Change: Assessing the Evidence, Geneva, Switzerland, pp. 353-384. Maslin, M. and P. Austin, 2012: Uncertainty: Climate models at their limit? Nature, 486(7402), 183. Masson-Delmotte, V., M. Schulz, A. Abe-Ouchi, J. Beer, A. Ganopolski, J.F.G. Rouco, E. Jansen, K. Lambeck, J. Luterbacher, T. Naish, T. Osborn, B. Otto-Bliesner, T. Quinn, R. Ramesh, M. Rojas, X. Shao, and A. Timmermann, 2014: Information from Paleoclimate Archives. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Mastrandrea, M.D., C.B. Field, T.F. Stocker, O. Edenhofer, K.L. Ebi, D.J. Frame, H. Held, E. Kriegler, K.J. Mach, P.R. Matschoss, G.K. Plattner, G.W. Yohe, and F.W. Zwiers, 2011: The IPCC AR5 guidance note on consistent treatment of uncertainties: a common approach across the working groups. Climatic Change, 108(4), 675. Maurer, E.P., L. Brekke, T. Pruitt, and P.B. Duffy, 2007: Fine-resolution climate projections enhance regional climate change impact studies. Eos, Transactions American Geophysical Union, 88(47), 504. Mawdsley, J.R., R. O MALLEY, and D.S. Ojima, 2009: A review of climate change adaptation strategies for wildlife management and biodiversity conservation. Conservation Biology, 23(5), 1080-1089. McCarthy, M.P., M. Best, and R. Betts, 2010: Climate change in cities due to global warming and urban effects. Geophysical Research Letters, 37(9). McGray, H., A. Hammill, R. Bradley, E.L. Schipper, and J.-. Parry, 2007: Weathering the Storm: Options for Framing Adaptation and Development. World Resources Institute, Washington, D.C., pp. 66. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McLeman, R.A., 2011: Settlement abandonment in the context of global environmental change. Global Environmental Change, Volume 21 (Suppl 1), S108-S120. McPhaden, M.J., S.E. Zebiak, and M.H. Glantz, 2006: ENSO as an Integrating Concept in Earth Science. Science, 314(5806), 1740-1745. McSweeney, C.F., R.G. Jones, and B.B. Booth, 2012: Selecting ensemble members to provide regional climate change information. Journal of Climate, 25(2012), 7100-7121. Mearns, L.O., M. Hulme, T.R. Carter, R. Leemans, M. Lal, and P. Whetton, 2001: Climate Scenario Development (Chapter 13). In: Climate Change 2001: The Scientific Basis, Contribution of Working Group I to the Third Assessment Report of the IPCC [Houghton, J.T. and et al. (eds.)]. Cambridge U. Press, Cambridge, pp. 583. Mearns, L.O., W.J. Gutowski, R. Jones, L.-. Leung, S. McGinnis, A.M.B. Nunes, and Y. Qian, 2009: A regional climate change assessment program for North America.. Eos, 90, 311. Mearns, L.O., S. Sain, R.L. Leung, M.S. Bukovsky, S. McGinnis, S. Biner, D. Caya, R.W. Arritt, W. Gutowski, E. Takle, M. Snyder, R.G. Jones, A.M.B. Nunes, S. Tucker, D. Herzmann, L. McDaniel, and L. Sloan, 2013: Climate change projections of the North American Regional Climate Change Assessment Program (NARCCAP). Climatic Change Letters, DOI:10.1007/s10584-013-0831-3. Mearns, L.O., R. Arritt, S. Biner, M.S. Bukovsky, S. McGinnis, S. Sain, D. Caya, J. Correia Jr, D. Flory, and W. Gutowski, 2012: The North American regional climate change assessment program: Overview of phase I results. Bulletin of the American Meteorological Society, 93(9), 1337-1362. Medlyn, B.E., R.A. Duursma, and M.J.B. Zeppel, 2011: Forest productivity under climate change: a checklist for evaluating model studies. Wiley Interdisciplinary Reviews: Climate Change, 2, 332-355. Mehta, V., G. Meehl, L. Goddard, J. Knight, A. Kumar, M. Latif, T. Lee, A. Rosati, and D. Stammer, 2011: Decadal climate predictability and prediction: Where are we? Bulletin of the American Meteorological Society, 92(5), 637-640. Meinshausen, M., S.J. Smith, K. Calvin, J.S. Daniel, M.L. Kainuma, J. Lamarque, K. Matsumoto, S. Montzka, S.C. Raper, and K. Riahi, 2011: The RCP greenhouse gas concentrations and their extensions from 1765 to 2300. Climatic Change, 109(1), 213-241. Meleux, F., F. Solmon, and F. Giorgi, 2007: Increase in European summer ozone amounts due to climate change. Atmospheric Environment, 41, 7577. Mendler de Suarez, J., P. Suarez, C. Bachofen, N. Fortugno, J. Goentzel, P. Gonc alves, N. Grist, C. Macklin, K. Pfeifer, S. Schweizer, M. Van Aalst, and H. Virji, 2012: Games for a New Climate: Experiencing the Complexity of Future Risks. . In: Pardee Center Taks Force Reports. Boston University, Boston, MA, pp. 109. Menéndez, C.G., M. de Castro, A. Sorensson, and J. Boulanger, 2010: CLARIS Project: towards climate downscaling in South America. Meteorologische Zeitschrift, 19(4), 357-362. Mesinger, F., G. DiMego, E. Kalnay, K. Mitchell, P.C. Shafran, W. Ebisuzaki, D. Jovic, J. Woollen, E. Rogers, and E.H. Berbery, 2006: North American regional reanalysis. Bulletin of the American Meteorological Society, 87(3), 343-360. Mickwitz, P., F. Aix, S. Beck, D. Carss, N. Ferrand, G. Görg, A. Jensen, P. Kivimaa, C. Kuhlicke, W. Kuindersma, M. Mánez, M. Melanen, S. Monni, A.B. Pedersen, R. H., and S. van Bommel, 2009: Climate Policy Integration, Coherence and Governance. In: PEER Report No 2. Partnership for European Environmental Research, Helsinki, Finland, pp. 92. Miles, E.L., M.M. Elsner, J.S. Littell, L.W. Binder, and D.P. Lettenmaier, 2010: Assessing regional impacts and adaptation strategies for climate change: the Washington Climate Change Impacts Assessment. Climatic Change, 102(1), 9-27. Morice, C.P., J.J. Kennedy, N.A. Rayner, and P.D. Jones, 2012: Quantifying uncertainties in global and regional temperature change using an ensemble of observational estimates: The HadCRUT4 data set. Journal of Geophysical Research, 117(D8), D08101. Morse, A., C. Prentice, and T. and Carter, 2009: Assessments of climate change impacts. In: ENSEMBLES: Climate Change and Its Impacts. [P. van der Linden and J. Mitchell (ed.)]. Met. Office Hadley Center, Exeter, UK, pp. 107. Moss, R.H., A.L. Brenkert, and E.L. Malone, 2001: Vulnerability to Climatic Change: An Quantitative Approach. PNNL-SA-33642. Pacific Northwest National Laboratory, Washington, D.C. Moss, R.H., J.A. Edmonds, K. Hibbard, M. Manning, S.K. Rose, D.P. van Vuuren, T.R. Carter, S. Emori, M. Kainuma, T. Kram, G. Meehl, J. Mitchell, N. Nakicenovic, K. Riahi, S. Smith, R.J. Stouffer, A. Thomson, J. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Weyant, and T. Wilbanks, 2010: The next generation of scenarios for climate change research and assessment. Nature, 463, 747-756. Mueller, J.M. and J.J. Hellmann, 2008: An assessment of invasion risk from assisted migration. Conservation Biology, 22, 562-567. Mukheibir, P., 2007: Possible climate change impacts on large hydroelectricity schemes in Southern Africa. Journal of Energy in Southern Africa, 18, 4. Murazaki, K. and P. Hess, 2006: How does climate change contribute to surface ozone change over the United States? . Journal of Geophysical Research, 111. Murphy, J., D. Sexton, G. Jenkins, and P. Boorman, 2009: Online Climate Change Projections Report. UKCP09 Climate Change Projections. Available at: http://ukclimateprojections.defra.gov.uk/content/view/824/517. Myers, N., 1988: Threatened biotas: "hotspots" in tropical forests. Environmentalist, 8, 187-208. Naik, A., 2009: Migration and Natural Disasters.[Laczko, F. and C. Aghazarm(eds.)]. Proceedings of Migration, Environment and Climate Change: Assessing the Evidence, Geneva, Switzerland, pp. 245-317. Nawaz, R., T. Bellerby, and M. Elshamy, Quantifying uncertainties in the assessment of Blue Nile flow sensitivity to climate change. Hydrology Sciences Journal, . Nelson, G., A. Palazzo, C. Ringler, T. Sulser, and M. Batka, 2009: The Role of International Trade in Climate Change Adaptation. ICTSD IPC Platform on Climate Change, Agriculture and Trade, Issue Paper No. 4, International Centre for Trade and Sustainable Development, Geneva, Switzerland and International Food & Agricultural Trade Policy Council, Washington DC, USA, pp. 16. Nelson, G.C., M.W. Rosegrant, J. Koo, R. Robertson, T. Sulser, T. Zhu, C. Ringler, S. Msangi, A. Palazzo, M. Batka, M. Magalhaes, R. Valmonte-Santos, M. Ewing, and D. Lee, 2009: Climate Change: Impact on Agriculture and Costs of Adaptation. International Food Policy Research Institute, Washington, D.C., pp. 19. New, M., A. Lopez, S. Dessai, and R. Wilby, 2007: Challenges in using probabilistic climate change information for impact assessments: an example from the water sector. Philosophical Transactions of the Royal Society A, 365, 2117-2131. Nicholls, R.J., S.E. Hanson, J.A. Lowe, R.A. Warrick, X. Lu, and A.J. Long, 2013: Sea-level scenarios for evaluating coastal impacts. Wiley Interdisciplinary Reviews: Climate Change, In press. Nóbrega, M., W. Collischonn, C. Tucci, and A. Paz, 2011: Uncertainty in climate change impacts on water resources in the Rio Grande Basin, Brazil. Hydrology and Earth System Sciences, 15(2), 585-595. Nordas, R. and N.P. Gleditsch, 2007: Climate change and conflict. Political Geography, 26, 627-638. Nunez, M.N., S.A. Solman, and M.F. Cabre, 2009: Regional climate change experiments over southern South America. II: Climate change scenarios in the late twentyfirst century. Climate Dynamics, 32, 1082. Nunez, M.N. and J.L. McGregor, 2007: Modeling future water environments of Tasmania, Australia. Climate Research, 34, 25. O Neill, B.C. and V. Schweizer, 2011: O Neill, B.C. and V. Schweizer, 2011: Projection and prediction: Mapping the road ahead. Nature Climate Change, 1, 352-353. Nature Climate Change, 1, 352-353. Ogallo, L., P. Bessemoulin, J. Ceron, S. Mason, and S.J. Connor, 2008: Adapting to climate variability and change: the Climate Outlook Forum process. Bulletin of the World Meteorological Organization, 57(2), 93-102. Ogden, N.H., L. St-Onge, I.K. Barker, S. Brazeau, M. Bigras-Poulin, D.F. Charron, C.M. Francis, A. Heagy, L.R. Lindsay, A. Maarouf, P. Michel, F. Milord, C.J. O'Callaghan, L. Trudel1, and R.A. Thompson, 2008: Risk maps for range expansion of the Lyme disease vector, Ixodes scapularis, in Canada now and with climate change. International Journal of Health Geographics, 7. Oh, C.H. and R. Reuveny, 2010: Climatic natural disasters, political risk, and international trade. Global Environmental Change, 20, 243-254. Oliver-Smith, A., 2012: Debating environmental migration: society, nature and population displacement in climate change. Journal of International Development, 24, 1058-1070. Olsson, J., W. Wang, L.P. Graham, J. Rosberg, and J. Andreasson, 2011: Using an ensemble of climate projections for simulating recent and near-future hydrological change to lake Vanern in Sweden. Tellus Series a-Dynamic Meteorology and Oceanography, 63A, 126. O'Neill, B.C. and V. Schweizer, 2011: Projection and prediction: Mapping the road ahead. Nature Climate Change, 1(7), 352-353. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 O'Neill, B.C., E. Kriegler, K. Riahi, K.L. Ebi, S. Hallegatte, T.R. Carter, R. Mathur, and D.P. van Vuuren, 2013: A new scenario framework for climate change research: The concept of Shared Socio-economic Pathways. Climatic Change, 10.1007/s10584-013-0905-2. Osborne, T., G. Rose, and T. Wheeler, 2013: Variation in the global-scale impacts of climate change on crop productivity due to climate model uncertainty and adaptation. Agricultural and Forest Meteorology, 170, 183- 194. Paeth, H. and M. Diederich, 2010: Postprocessing of simulated precipitation for impact studies in West Africa. Part II: A weather generator for daily data. Climate Dynamics, 36(7-8), 1337. Paeth, H., N.M.J. Hall, M.A. Gaertner, M.D. Alonso, S. Moumouni, J. Polcher, P.M. Ruti, A.H. Fink, M. Gosset, T. Lebel, A.T. Gaye, D.P. Rowell, W. Moufouma-Okia, D. Jacob, B. Rockel, F. Giorgi, and M. Rummukainen, 2011: Progress in regional downscaling of west African precipitation. Atmospheric Science Letters, 12(1), 75. Palosuo, T., K.C. Kersebaum, C. Angulo, P. Hlavinka, M. Moriondo, J.E. Olesen, R.H. Patil, F. Ruget, C. Rumbaur, J. Takáè, M. Trnka, M. Bindi, B. Çalda , F. Ewert, R. Ferrise, W. Mirschel, L. ªaylan, B. ©i¹ka, and R. Rötter, 2011: Simulation of winter wheat yield and its variability in different climates of Europe: A comparison of eight crop growth models. European Journal of Agronomy, 35, 103-114. Parmesan, C., C. Duarte, E. Poloczanska, A.J. Richardson, and M.C. Singer, 2011: Overstretching attribution. Nature Climate Change, 1, 2-4. Parry, M., N. Arnell, P. Berry, D. Dodman, S. Fankhauser, C. Hope, S. Kovats, R. Nicholls, D. Satterthwaite, R. Tiffin, and T. Wheeler, 2009: Assessing the Costs of Adaptation to Climate Change: A Review of the UNFCCC and Other Recent Estimates. International Institute for Environment and Development and Grantham Institute for Climate Change, Imperial College London, UK, pp. 111. Patt, A.G., D.P. van Vuuren, F. Berkhout, A. Aaheim, A.F. Hof, M. Isaac, and R. Mechler, 2010: Adaptation in integrated assessment modeling: where do we stand? Climatic Change, 99, 383-402. Perch-Nielsen, S.L., M.B. Bättig, and D. Imboden, 2008: Exploring the link between climate change and migration. Climatic Change, 91, 375-393. Perrels, A., N. Veijalainen, J. Jylhä, J. Aaltonen, R. Molarius, M. Porthin, J. Silander, T. Rosqvist, T. Tuovinen, T. Carter, and S. Fronzek, 2010: The implications of climate change for extreme weather events and their socio- economic consequences in Finland. In: VATT Research Reports 158/2010. Government Institute for Economic Research, Helsinki, pp. 133. Peters, G.P., R.M. Andrew, T. Boden, J.G. Canadell, P. Ciais, C. Le Quéré, G. Marland, M.R. Raupach, and C. Wilson, 2013: The challenge to keep global warming below 2 °C. Nature Climate Change, 3(1), 4. Peters, G.P., 2010: Managing carbon leakage. Carbon Management, 1(1), 35-37. Peters, G.P., J.C. Minx, C.L. Weber, and O. Edenhofer, 2011: Growth in emission transfers via international trade from 1990 to 2008. Proceedings of the National Academy of Science, 108(21), 8903-8908. Piguet, E., 2010: Linking climate change, environmental degradation, and migration: a methodological overview. Wiley Interdisciplinary Reviews: Climate Change, 1, 517-524. Preston, B.L., E.J. Yuen, and R.M. Westaway, 2011: Putting vulnerability to climate change on the map: a review of approaches, benefits, and risks. Sustainability Science, 6, 177 202. PriceWaterhouseCoopers, 2010: Adapting to climate change in the infrastructure sectors. Maintaining robust and resilient infrastructure systems in the energy, transport, water and ICT sectors. A report for the Adapting to Climate Change Programme, UK Department for Environment, Food and Rural Affairs. . PriceWaterhouseCoopers, London, UK, pp. 100. Pryor, S. and R. Barthelmie, 2011: Assessing climate change impacts on the near-term stability of the wind energy resource over the United States. Proceedings of the National Academy of Sciences, 108(20), 8167-8171. Rauscher, S.A., F. Giorgi, N.S. Diffenbaugh, and A. Seth, 2008: Extension and intensification of the meso-american mid summer drought in the 21st century. Climate Dynamics, 31, 551. Rauscher, S.A., F. Kucharski, and A. Seth, 2011: The role of regional SST warming variations in the drying of Meso-America in future climate projections. Journal of Climate, 24, 2003. Rayner, N., D. Parker, E. Horton, C. Folland, L. Alexander, D. Rowell, E. Kent, and A. Kaplan, 2003: Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century. Journal of Geophysical Research, 108(D14), 4407. Reeder, T. and N. Ranger, 2010: How Do You Adapt in an Uncertain World? Lessons from the Thames Estuary 2100 Project. In: Expert Perspectives Series. World Resources Institute, Washington, DC. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rhein, M., S.R. Rintoul, S. Aoki, E. Campos, D. Chambers, R.A. Feely, S. Gulev, G.C. Johnson, S.A. Josey, A. Kostianoy, C. Mauritzen, D. Roemmich, L. Talley, and F. Wang, 2014: Observations: Ocean. In: Climate Change 2014: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Stocker, T.,et al. (ed.)]. pp. 12. Rienecker, M.M., M.J. Suarez, R. Gelaro, R. Todling, J. Bacmeister, E. Liu, M.G. Bosilovich, S.D. Schubert, L. Takacs, and G. Kim, 2011: MERRA: NASA's modern-era retrospective analysis for research and applications. Journal of Climate, 24(14), 3624-3648. Rogelj, J., M. Meinshausen, and R. Knutti, 2012: Global warming under old and new scenarios using IPCC climate sensitivity range estimates. Nature Climate Change, 2, 248. Rosenberg, E.A., P.W. Keys, D.B. Booth, D. Hartley, J. Burkey, A.C. Steinemann, and D.P. Lettenmaier, 2010: Precipitation extremes and the impacts of climate change on stormwater infrastructure in Washington State. Climatic Change, 102(1), 319-349. Rosenzweig, C., W. Solecki, L. Parshall, B. Lynn, J. Cox, and R. Goldberg, 2009: Mitigating New York City s heat island. Bulletin of the Amer. Met. Soc, 90, 1298. Rosenzweig, C., D. Karoly, M. Vicarelli, P. Neofotis, Q. Wu, G. Casassa, A. Menzel, T.L. Root, N. Estrella, B. Seguin, P. Tryjanowski, C. Liu, S. Rawlins, and A. Imeson, 2008: Attributing physical and biological impacts to anthropogenic climate change. Nature, 453, 353-358. Rosenzweig, C., W.D. Solecki, R. Blake, M. Bowman, C. Faris, V. Gornitz, R. Horton, K. Jacob, A. LeBlanc, and R. Leichenko, 2011: Developing coastal adaptation to climate change in the New York City infrastructure-shed: process, approach, tools, and strategies. Climatic Change, 106(1), 93-127. Rosenzweig, C., J.W. Jones, J.L. Hatfield, A.C. Ruane, K.J. Boote, P. Thorburn, J.M. Antle, G.C. Nelson, C. Porter, S. Janssen, S. Asseng, B. Basso, F. Ewert, D. Wallach, G. Baigorria, and J.M. Winter, 2013: The Agricultural Model Intercomparison and Improvement Project (AgMIP): Protocols and pilot studies. Agricultural and Forest Meteorology, 170, 166-182. Rosenzweig, C. and P. Neofotis, 2013: Detection and attribution of anthropogenic climate change impacts. Wiley Interdisciplinary Reviews: Climate Change, 4, 121-150. Roshydromet, 2008: Assessment Report on Climate Change and its Consequences in Russian Federation: General Summary. Federal Service for Hydrometeorology and Environmental Monitoring (Roshydromet), Moscow, pp. 24. Rötter, R.P., T.R. Carter, J.E. Olesen, and J.R. Porter, 2011: Crop-climate models need an overhaul. Nature Climate Change, 1, 175-177. Rounsevell, M.D.A. and M.J. Metzger, 2010: Developing qualitative scenario storylines for environmental change assessment. Wiley Interdisciplinary Reviews: Climate Change, 1, 606-619. Rowell, D.P. and R.G. Jones, 2006: Causes and uncertainty of future summer drying over Europe. Climate Dynamics, 27(2), 281-299. Rudolf, B., A. Becker, U. Schneider, A. Meyer-Christoffer, and M. Ziese, 2011: New GPCC full data reanalysis version 5 provides high-quality gridded monthly precipitation data. GEWEX News, 21(2), 4-5. Rygel, L., D. O Sullivan, and B. Yarnal, 2006: A method for constructing a social vulnerability index: an application to hurricane storm surges in a developed country. Mitigation and Adaptation Strategies for Global Change, 11(3), 741. Saha, S., S. Moorthi, H. Pan, X. Wu, J. Wang, S. Nadiga, P. Tripp, R. Kistler, J. Woollen, and D. Behringer, 2010: The NCEP climate forecast system reanalysis. Bulletin of the American Meteorological Society, 91(8), 1015- 1057. Salathé, E.P., R. Steed, C.F. Mass, and P.H. Zahn, 2008: A High-Resolution Climate Model for the U.S. Pacific Northwest: Mesoscale Feedbacks and Local Responses to Climate Change. Journal of Climate, 21, 5708-5726. Salathé, E.P., L.R. Leung, Y. Qian, and Y. Zhang, 2010: Regional climate model projections for the State of Washington. Climatic Change, 102(1), 51-75. Schär, C., P.L. Vidale, D. Lüthi, C. Frei, C. Häberli, M.A. Liniger, and C. Appenzeller, 2004: The role of increasing temperature variability in European summer heat wave. Nature, 427, 332. Schiermeier, Q., 2012: Models hone picture of climate impacts. Nature, 482, 286. Schipper, L., W. Liu, D. Krawanchild, and S. Chanthy, 2010: Review of climate change adaptation methods and tools. MRC Technical Paper no. 34. . Mekong River Commission, Vientiane, pp. http://www.mrcmekong.org/assets/Publications/technical/Tech-No34-Review-of-climate-change.pdf. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Schipper, L. and M. Pelling, 2006: Disaster risk, climate change and international development: scope for, and challenges to, integration. Disasters, 30, 19-38. Schoof, J.T., S.C. Pryor, and J. Suprenant, 2010: Development of daily precipitation projections for the United States based on probabilistic downscaling. Journal of Geophysical Reseaarch, 115(D13). Schweiger, O., R.K. Heikkinen, A. Harpke, T. Hickler, S. Klotz, O. Kudrna, I. Kühn, J. Pöyry, and J. Settele, 2012: Increasing range mismatching of interacting species under global change is related to their ecological characteristics. Global Ecology and Biogeography, 607, 88-99. Schweizer, V.J. and E. Kriegler, 2012: Improving environmental change research with systematic techniques for qualitative scenarios. Environmental Research Letters, 7, 044011. Searchinger, T., R. Heimlich, R.A. Houghton, F. Dong, A. Elobeid, J. Fabiosa, S. Tokgoz, D. Hayes, and T.-. Yu, 2008: Use of U.S. croplands for biofuels increases greenhouse gases through emissions from land-use change. Science, 319, 1238-1240. SEI, 2009: Economics of Climate Change: Kenya. Stockholm Environment Institute, Oxford Office, UK, pp. 67. Seneviratne, S.I., N. Nicholls, D. Easterling, C.M. Goodess, S. Kanae, J. Kossin, Y. Luo, J. Marengo, K. McInnes, M. Rahimi, M. Reichstein, A. Sorteberg, C. Vera, and X. Zhang, 2012: Changes in climate extremes and their impacts on the natural physical environment . In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. Cambridge University Press, Cambridge, United Kingdom, and New York, NY, USA, pp. 109-230. Seneviratne, S.I., D. Luthi, M. Litschi, and C. Schar, 2006: Investigating soil moisture-climate interactions in a changing climate : A review. Nature, 443, 205. Shindell, D.T., G. Faluvegi, D.S. Stevenson, M.C. Krol, L.K. Emmons, J.-. Lamarque, G. Pétron, F.J. Dentener, K. Ellingsen, M.G. Schultz, O. Wild, M. Amann, C.S. Atherton, D.J. Bergmann, I. Bey, T. Butler, J. Cofala, W.J. Collins, R.G. Derwent, R.M. Doherty, J. Drevet, and et al., 2006: Multimodel simulations of carbon monoxide: Comparison with observations and projected near-future changes. Journal of Geophysical Research: Atmospheres, 111(D19). Shongwe, M.E., G.J. van Oldenborgh, B.J.J.M. van den Hurk, B. de Boer, C.A.S. Coelho, and M.K. van Aalst, 2009: Projected Changes in Mean and Extreme Precipitation in Africa under Global Warming. Part I: Southern Africa. Journal of Climate, 22(13), 3819-3837. Shrestha, R.R., Y.B. Dibike, and T.D. Prowse, 2011: Modelling of climate-induced hydrologic changes in the Lake Winnipeg watershed. Journal of Great Lakes Research, . Sietz, D., M.K.B. Lu deke, and C. Walther, 2011: Categorisation of typical vulnerability patterns in global drylands. . Global Environmental Change, 21, 431-440. Sitch, S., C. Huntingford, N. Gednev, P.E. Levy, M. Lomas, S.L. Piao, R. Betts, P. Ciais, P. Cox, P. Friedlingstein, C.D. Jones, I.C. Prentice, and F.I. Woodward, 2008: Evaluation of the terrestrial carbon cycle, future plant geography and climate-carbon cycle feedbacks using five Dynamic Global Vegetation Models (DGVMs). Global Change Biology, 14, 1-25. Smit, B. and J. Wandel, 2006: Adaptation, adaptive capacity and vulnerability. Global Environmental Change, 16(3), 282-292. Smith, A., 2009: Global Warming Reopens the Northeast Passage. Time Magazine, 17/09/09. Smith, L.C. and S.R. Stephenson, 2013: New Trans-Arctic shipping routes navigable by midcentury. Proceedings of the National Academy of Sciences, . Somanathan, E., T. Sterner, T. Sugiyama, D. Chimanikire, N. Dubash, J.K. Essandoh-Yeddu, S. Fifita, L. Goulder, A. Jaffe, X. Labandeira, S. Managi, C. Mitchell, J.P. Montero, F. Teng, and T. Zylicz, 2014: National and sub- national policies and institutions. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O.,et al. (ed.)]. Sommer, U., N. Aberle, A. Engel, T. Hansen, K. Lengfellner, M. Sandow, J. Wohlers, E. Zöllner, and U. Riebesell, 2007: An indoor mesocosm system to study the effect of climate change on the late winter and spring succession of Baltic Sea phyto- and zooplankton. Oecologia, 150(4), 655-667. Song, R.Y., X.J. Gao, H.Q. Zhang, and A. Moise, 2008: 20 km resolution regional climate model experiments over Australia: experimental design and simulations of current climate. Australian Meteorological Magazine, 57, 175. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sorensson, A.A., C.G. Menendez, R. Ruscica, P. Alexander, P. Samuelsson, and U. Willen, 2010: Projected precipitatiaon changes in South America: A dynamical downscaling within CLARIS. Meteorologische Zeitschrift, 19, 347. Stafford Smith, M., L. Horrocks, A. Harvey, and C. Hamilton, 2011: Rethinking adaptation for a 4°C world. Phil. Trans. R. Soc. A, 369, 196-216. Stavins, R., Z. Ji, T. Brewer, M. Conte Grand, M. den Elzen, M. Finus, J. Gupta, N. Höhne, M.-. Lee, A. Michaelowa, M. Paterson, K. Ramakrishna, G. Wen, J. Wiener, and H. Winkler, 2014: International cooperation: agreements and instruments. In: Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. [Edenhofer, O.,et al. (ed.)]. Steiner, A.L., S. Tonse, R.C. Cohen, A.H. Goldstein, and R.A. Harley, 2006: Influence of future climate and emissions on regional air quality in California. Journal of Geophysical Research, 111(D18). Stephenson, S.R., L.C. Smith, and J.A. Agnew, 2011: Divergent long-term trajectories of human access to the Arctic. Nature Climate Change, 1, 156-160. Stephenson, S.R., L.W. Brigham, L.C. Smith, and J.A. Agnew, 2013: Projected 21st-century changes to Arctic marine access. Climatic Change, 118(3-4), 885-899. Stern, N., 2007: The Economics of Climate Change: The Stern Review. Cambridge University Press, Cambridge, UK, pp. 692. Stevenson, D.S., F.J. Dentener, M.G. Schultz, K. Ellingsen, T.P.C. van Noije, O. Wild, G. Zeng, and et al., 2006: Multimodel ensemble simulations of present day and near future tropospheric ozone. Journal of Geophysical Research, 111(D8). Stokke, O.S., 2007: A legal regime for the Arctic? Interplay with the Law of the Sea Convention. Marine Policy, 31(4), 402-408. Stokke, O.S., 2007: A legal regime for the Arctic? Interplay with the Law of the Sea Convention. Marine Policy, 31(4), 402-408. Stokke, O.S., 2007: A legal regime for the Arctic? Interplay with the Law of the Sea Convention. Marine Policy, 31(4), 402-408. Stokke, O.S., 2007: A legal regime for the Arctic? Interplay with the Law of the Sea Convention. Marine Policy, 31(4), 402-408. Subin, Z.M., W.J. Riley, J. Jin, D.S. Christianson, M.S. Torn, and L.M. Kueppers, 2011: Ecosystem Feedbacks to Climate Change in California: Development, Testing, and Analysis Using a Coupled Regional Atmosphere and Land Surface Model (WRF3 CLM3.5). Earth Interactions, 15, 1. Sutherst, R.W., F. Constable, K.J. Finlay, R. Harrington, J. Luck, and M.P. Zalucki, 2011: Adapting to crop pest and pathogen risks under a changing climate. Wiley Interdisciplinary Reviews: Climate Change, 2, 220-237. Swanson, D.A., J.C. Hiley, H.D. Venema, and R. Grosshans, 2007: Indicators of adaptive capacity to climate change for agriculture in the prairie region of Canada: an analysis based on Statistics Canada s Census of Agriculture. Working Paper for the Prairie Climate Resilience Project. International Institute for Sustainable Development (IISD), Winnipeg, Ontario, Canada. Szopa, S. and D.A. Hauglustaine, 2007: Relative impacts of worldwide tropospheric ozone changes and regional emission modifications on European surface ozone levels. Comptes Rendus Geoscience, 339(11-12), 709. Tänzler, D., A. Maas, and A. Carius, 2010: Climate change adaptation and peace. Wiley Interdisciplinary Reviews: Climate Change, 1, 741 750. Tao, F. and Z. Zhang, 2013: Climate change, wheat productivity and water use in the North China Plain: A new super-ensemble-based probabilistic projection. Agricultural and Forest Meteorology, 170, 146-165. Taylor, C.M., D.J. Parker, N. Kalthoff, M.A. Gaertner, N. Philippon, S. Bastin, P.P. Harris, A. Boone, and et al., 2011: New perspectives on land-atmosphere feedbacks from the African Monsoon Multidisciplinary Analysis. Atmospheric Science Letters, 12(1), 38. Taylor, K.E., R.J. Stouffer, and G.A. Meehl, 2012: A summary of the CMIP5 experiment design. Bulletin of the American Meteorological Society, 93, 485-498. Tebaldi, C. and D. Lobell, 2008: Towards probabilistic projections of climate change impacts on global crop yields. Geophysical Research Letters, 35(8), L08705. Tebaldi, C., J.M. Arblaster, and R. Knutti, 2011: Mapping model agreement on future climate projections. Geophysical Research Letters, 38(23). Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Thomas, C.D., 2011: Translocation of species, climate change, and the end of trying to recreate past ecological communities. Trends in Ecology & Evolution, 26, 216-221. Thorne, P.W. and R.S. Vose, 2010: Reanalyses suitable for characterizing long-term trends are they really achievable? Bulletin of the American Meteorological Society, 91, 353. Timbal, B., P. Hope, and S. Charles, 2008: Evaluating the consistency between statistically downscaled and global dynamical model climate change projections. Journal of Climate, 21, 6052. Todd, M.C., R.G. Taylor, T.J. Osborn, D.G. Kingston, N.W. Arnell, and S.N. Gosling, 2011: Uncertainty in climate change impacts on basin-scale freshwater resources preface to the special issue: the QUEST-GSI methodology and synthesis of results. Hydrology and Earth System Sciences, 15, 1035-1046. Tolika, C.K., P. Zanis, and C. Anagnostopoulu, 2012: Regional climate change scenarios for Greece: future temperature and precipitation projections from ensembles of RCMs. Global NEST Journal, 14(4), 407-421 (Greece, regional climate models, emission scenarios, precipitation and temperature future changes.). Traill, L.W., M.L.M. Lim, N.S. Sodhi, and C.J.A. Bradshaw, 2010: Mechanisms driving change: altered species interactions and ecosystem function through global warming. Journal of Animal Ecology, 79, 937-947. Tran, P., R. Shaw, G. Chantry, and J. Norton, 2009: GIS and local knowledge in disaster management: a case study of flood risk mapping in Viet Nam. Disasters, 33(1), 152-169. Trenberth, K.E., A. Dai, R.M. Rasmussen, and D.B. Parsons, 2003: Bulletin of the American Meteorological Society, 84(9), 1205-1217. Trenberth, K.E., 2011: Changes in precipitation with climate change. Climate Research, 47, 123. Trivedi, M.R., P.M. Berry, M.D. Morecroft, and T.P. Dawson, 2008: Spatial scale affects bioclimate model projections of climate change impacts on mountain plants. Global Change Biology, 14, 1089-1103. Trostle, R., D. Marti, S. Rosen, and P. Westcott, 2011: Why Have Food Commodity Prices Risen Again?. Economic Research Service, United States Department of Agriculture, pp. 29. Tzanopoulos, J., R. Mouttet, A. Letourneau, I.N. Vogiatzakis, S.G. Potts, K. Henle, R. Mathevet, and P. Marty, 2013: Scale sensitivity of drivers of environmental change across Europe. Global Environmental Change, 23, 167-178. UNFCCC, 2007: Investment and Financial Flows to Address Climate Change. United Nations Framework Convention on Climate Change Secretariat, Bonn, Germany, pp. 272. UNFCCC, 2008: Compendium on Methods and Tools to Evaluate Impacts of and Vulnerability to Climate Change. UNFCCC, Bonn, Germany. UNFPA, 2007: State of the World Population 2007. United Nations Population Fund, New York, pp. 100. van Aalst, M.K., T. Cannon, and I. Burton, 2008: Community level adaptation to climate change: The potential role of participatory community risk assessment. Global Environmental Change-Human and Policy Dimensions, 18(1), 165-179. van Ruijven, B.J., M. Levy, A. Agrawal, F. Biermann, J. Birkmann, T.R. Carter, K.L. Ebi, M. Garschagen, B. Jones, R. Jones, E. Kemp Benedict, M. Kok, K. Kok, M.C. Lemos, P.L. Lucas, B. Orlove, S. Pachauri, T. Parris, A. Patwardhan, A. Petersen, B.L. Preston, J. Ribot, D.S. Rothman, and V.J. Schweizer, 2013: Enhancing the relevance of Shared Socioeconomic Pathways for climate change impacts, adaptation and vulnerability research. Climatic Change, doi: 10.1007/s10584-013-0931-0. Van Vuuren, D.P. and B.C. O'Neill, 2006: The consistency of IPCC's SRES scenarios to 1990 2000 trends and recent projections. Climatic Change, 75(1), 9-46. van Vuuren, D.P., S.J. Smith, and K. Riahi, 2010: Downscaling socioeconomic and emissions scenarios for global environmental change research: a review. WIREs Climate Change, 1(3), 393-404. van Vuuren, D.P., J. Edmonds, M. Kainuma, K. Riahi, A. Thomson, K. Hibbard, G.C. Hurtt, T. Kram, V. Krey, J.-. Lamarque, T. Masui, M. Meinshausen, N. Nakicenovic, S.J. Smith, and S.K. Rose, 2011: The representative concentration pathways: an overview. Climatic Change, 109(1-2), 5-31. van Vuuren, D.P., K. Riahi, R. Moss, J. Edmonds, A. Thomson, N. Nakicenovic, T. Kram, F. Berkhout, R. Swart, A. Janetos, S.K. Rose, and N. Arnell, 2012: A proposal for a new scenario framework to support research and assessment in different climate research communities. Global Environmental Change, 22, 21-35. van Vuuren, D.P., E. Kriegler, B.C. O Neill, K.L. Ebi, K. Riahi, T.R. Carter, J. Edmonds, S. Hallegatte, T. Kram, R. Mathur, and H. Winkler, 2013: A new scenario framework for climate change research: scenario matrix architecture. Climatic Change, doi: 10.1007/s10584-013-0906-1. Villa, F. and H. McLEOD, 2002: Environmental vulnerability indicators for environmental planning and decision- making: guidelines and applications. Environmental Management, 29(3), 335-348. Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wang, E., P. McIntosh, Q. Jiang, and J. Xu, 2009: Quantifying the value of historical climate knowledge and climate forecasts using agricultural systems modelling. Climatic Change, 96, 45-61. Ward, P., S. Pelt, O. Keizer, J. Aerts, J. Beersma, B. den Hurk, and A. te Linde, 2013: Including climate change projections in probabilistic flood risk assessment. Journal of Flood Risk Management, . Warner, K., 2010: Global environmental change and migration: Governance challenges. Global Environmental Change, 20, 402-413. Washington, R., M. Harrison, D. Conway, E. Black, A. Challinor, D. Grimes, R. Jones, A. Morse, G. Kay, and M. Todd, 2006: African Climate Change. Watkiss, P. and A. Hunt , 2012: Projection of economic impacts of climate change in sectors of Europe based on bottom up analysis: human health. Climatic Change, 112(1), 101. Watkiss, P., 2011: Aggregate economic measures of climate change damages: explaining the differences and implications. Wiley Interdisciplinary Reviews: Climate Change, 2, 356-372. Watkiss, P. (ed.), 2011: The ClimateCost Project. Final Report. Volume 1: Europe. Stockholm Environment Institute, Sweden, . Watterson, I.G., 2008: Calculation of probability density functions for temperature and precipitation change under global warming. Journal of Geophysical Research-Atmospheres, 113(D12106). Watterson, I.G., J.L. McGregor, and K.C. Nguyen, 2008: Changes in extreme temperatures in Australasian summer simulated by CCAM under global warming, and the role of winds and land-sea contrasts. Australian Meteorological Magazine, 57, 195. Watterson, I., 2011: Calculation of joint PDFs for climate change with properties matching recent Australian projections. Australian Meteorological and Oceanographic Journal, 61(4), 211. Watterson, I. and P. Whetton, 2011: Joint PDFs for Australian climate in future decades and an idealized application to wheat crop yield. Australian Meteorological and Oceanographic Journal, 61(4), 221. Weaver, C.P., E. Cooter, R. Gilliam, A. Gilliland, A. Grambsch, D. Grano, B. Hemming, S.W. Hunt, C. Nolte, and D.A. Winner, 2009: A preliminary synthesis of modeled climate change impacts on U.S. regional ozone concentrations. Bulletin of the American Meteorological Society, 90, 1843. Wetterhall, F., L. Graham, J. Andréasson, J. Rosberg, and W. Yang, 2011: Using ensemble climate projections to assess probabilistic hydrological change in the Nordic region. Nat.Hazards Earth Syst.Sci, 11, 2295-2306. Wetterhall, F., L.P. Graham, J. Andréasson, J. Rosberg, and W. Yang, 2011: Using ensemble climate projections to assess probabilistic hydrological change in the Nordic region. Natural Hazards and Earth System Science, 11, 2295-2306. Wetterhall, F., L.P. Graham, J. Andréasson, J. Rosberg, and W. Yang, 2011: Using ensemble climate projections to assess probabilistic hydrological change in the Nordic region. Natural Hazards and Earth System Science, 11, 2295-2306. Wetterhall, F., L.P. Graham, J. Andréasson, J. Rosberg, and W. Yang, 2011: Using ensemble climate projections to assess probabilistic hydrological change in the Nordic region. Natural Hazards and Earth System Science, 11, 2295-2306. Wetterhall, F., L.P. Graham, J. Andréasson, J. Rosberg, and W. Yang, 2011: Using ensemble climate projections to assess probabilistic hydrological change in the Nordic region. Natural Hazards and Earth System Science, 11, 2295-2306. White, J.W., G. Hoogenboom, B.A. Kimball, and G.W. Wall, 2011: Methodologies for simulating impacts of climate change on crop production. Field Crops Research, 124, 357-368. Wilby, R.L., J. Troni, Y. Biot, L. Tedd, B.C. Hewitson, D.M. Smith, and R.T. Sutton, 2009: A review of climate risk information for adaptation and development planning. International Journal of Climatology, 29, 1193-1215. Williams, A.P., C.D. Allen, A.K. Macalady, D. Griffin, C.A. Woodhouse, D.M. Meko, T.W. Swetnam, S.A. Rauscher, R. Seager, and H.D. Grissino-Mayer, 2012: Temperature as a potent driver of regional forest drought stress and tree mortality. Nature Climate Change, . Williams, J.W., S.T. Jackson, and J.E. Kutzbach, 2007: Projected distributions of novel and disappearing climates by 2100 AD. Proceedings of the National Academy of Science, 104, 5738-5742. Willows, R. and R. Connell, 2003: Climate Adaptation: Risk, Uncertainty, and Decision-making. UKCIP Technical Report, UK Climate Impacts Programme, Oxford, pp. 154. Wisner, B., P. Blaikie, T. Cannon, and I. Davis, 2005: At Risk, Natural Hazards, People s Vulnerability and Disasters . Routledge, London and New York, pp. 297. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 World Bank, 2010: Economics of Adaptation to Climate Change. Social Synthesis Report. . World Bank, Washington, DC, pp. 116. World Bank, 2010: The Costs to Developing Countries of Adapting to Climate Change: New Methods and Estimates. The Global Report of the Economics of Adaptation to Climate Change Study, Consultation Draft, Washington, D.C., pp. 84. World Weather Watch, 2005: World Weather Watch: Twenty-second Status Report on Implementation. Secretariat of the World Meteorological Organization, . WTO-UNEP, 2009: Trade and Climate Change. World Trade Organization and United Nations Environment Programme, Geneva, Switzerland, pp. 166. Xu, Y., X.J. Gao, and F. Giorgi, 2009: Regional variability of climate change Hotspots in East Asia. Advances in Atmospheric Sciences, 26, 783. Yatagai, A., K. Kamiguchi, O. Arakawa, A. Hamada, N. Yasutomi, and A. Kitoh, 2012: APHRODITE: Constructing a long-term daily gridded precipitation dataset for Asia based on a dense network of rain gauges. Bulletin of the American Meteorological Society, 93(9), 1401-1415. Yin, C., Y. Li, W. Ye, J.F. Bornman, and X. Yan, 2010: Statistical downscaling of regional daily precipitation over southeast Australia based on self-organizing maps. Theoretical and Applied Climatology, 105(1-2), 11. Zanis, P., I. Kapsomenakis, C. Philandras, K. Douvis, D. Nikolakis, E. Kanellopoulou, C. Zerefos, and C. Repapis, 2009: Analysis of an ensemble of present day and future regional climate simulations for Greece. International Journal of Climatology, 29(11), 1614-1633. Zanis, P., E. Katragkou, I. Tegoulias, A. Poupkou, D. Melas, P. Huszar, and F. Giorgi, 2011: Evaluation of near surface ozone in air quality simulations forced by a regional climate model over Europe for the period 1991 2000. Atmospheric Environment, 45(36), 6489-6500. Zhang, Q., H. Körnich, and K. Holmgren, 2012: How well do reanalyses represent the southern African precipitation? Climate Dynamics, 40(3-4), 951. Zhang, X., L. Alexander, G.C. Hegerl, P. Jones, A.K. Tank, T.C. Peterson, B. Trewin, and F.W. Zwiers, 2011: Indices for monitoring changes in extremes based on daily temperature and precipitation data. Wiley Interdisciplinary Reviews: Climate Change, . Zhang, Y., X.M. Hu, L.R. leung, and W.I. Gustafson, 2008: Impacts of regional climate change on biogenic emissions and air quality. Journal of Geophysical Research, 113. Ziervogel, G. and A. Taylor, 2008: Feeling stressed: integrating climate adaptation with other priorities in South Africa. Environment: Science and Policy for Sustainable Development, 50(2), 32-41. Ziervogel, G. and P.J. Ericksen, 2010: Adapting to climate change to sustain food security. Wiley Interdisciplinary Reviews: Climate Change, 1, 525-540. Ziervogel, G., P. Johnston, M. Matthew, and P. Mukheibir, 2010: Using climate information for supporting climate change adaptation in water resource management in South Africa. Climatic Change, 103, 537-554. Zoomers, A., 2011: Introduction: Rushing for Land: Equitable and sustainable development in Africa, Asia and Latin America. Development, 54(1), 12-20. Zurek, M.B. and T. Henrichs, 2007: Linking scenarios across geographical scales in international environmental assessments. . Technological Forecasting and Social Change, 74, 1282-1295. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-1: Dimensions of the institutions and actors involved in climate change decision-making, including example entries referred to in chapters of this volume. Vertical integration can occur within as well as between levels. Decison-making domains are illustrative. Modified and extended from Mickwitz (2009). Domain: Economy Energy Food/fibre Technology Environment ... Coherent policies and decision-making Level: Multi-level organisation and governance IMF/WB IEA FAO WIPO UNFCCC WTO NGOs WTO NGOs CBD Global MDGs CLOS  (fisheries) Montreal Protocol NGOs NGOs NGOs MFIs/MDBs OPEC AFTA Multi-­ nationals R&D CLRTAP Trans-­ BFIs Electric  grid   COMESA EU  Innovation  Union MRC national   OECD/EU operators MERCOSUR LVBC CLOS  (transport) Oil/gas  distributor EU CAP/CFP EU Directives Ministry/Gov.   Ministry/Gov.   Ministry/Gov.   Ministry/Gov.   Ministry/Gov.   Dept./Agency Dept./Agency Dept./Agency Dept./Agency   Dept./Agency National   Banks Energy  provider Tariffs, Quotas,   Education/R&D/ Environmental law Taxation Energy  regulator Regulations Innovation State/Province/ State/Province/ State/Province/ State/Province/ State/Province/ Sub-­ County/City   County/City   County/City   County/City   County/City   national   Taxation Public/private Extension service Incentives,  Science   Protected  areas energy  provider Land use  planning parks   Regional offices   Micro-­ finance, Renewables   Farmer, Forester,   Entrepreneur, Environmentalist, Co-­ operative,   Producer,    Voter,   Fisher,   Invester,  Voter, Landowner,  Voter,   Local Employer, Voter,   Consumer Landowner,   Consumer Consumer Consumer Voter,  Consumer Acronyms: IMF International Monetary Fund; WB World Bank; WTO World Trade Organization; MDGs Millennium Development Goals, NGO Non-governmental Organization; MDBs Multilateral Development Banks; MFIs Multilateral Financial Institutions; BFIs Bilateral Finance Institutions; ECD Organisation for Economic Co-operation and Development; EU European Union; CLOS United Nations Convention on the Law of the Sea; IEA International Energy Agency; OPEC Organization of the Petroleum Exporting Countries; FAO Food and Agriculture Organization of the United Nations; AFTA Association of Southeast Asian Nations (ASEAN) Free Trade Area; COMESA Common Market for Eastern and Southern Africa; MERCOSUR Mercado Común del Sur (Southern Common Market); CAP/CFP Common Agricultural Policy/Common Fisheries Policy; WIPO World Intellectual Property Organization, UNFCCC United Nations Framework Convention on Climate Change; CBD Convention on Biological Diversity; CLRTAP Convention on Long-range Transboundary Air Pollution (Europe, N. America, C. Asia); MRC Mekong River Commission For Sustainable Development; Lake Victoria Basin Commission. Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-2: Selected examples of regional treatment in previous IPCC Assessment Reports and Special Reports (SR). Major assessments are subdivided by the three Working Group reports, each described by generic titles. IPCC report Year Treatment of regions [references] First Assessment 1990 Climate: Climate projections for 2030 in 5 sub-continental regions; Observations averaged for Report (FAR) northern/southern hemisphere, by selected regions and by 20° latitude x 60° longitude grid boxes [1, 2, 3] Impacts: Agriculture by continent (7 regions); Ecosystem impacts for 4 biomes; water resources for case study regions; Oceans and Coastal Zones treated separately Responses: Emissions scenarios by 5 economic groupings; Energy and Industry by 9 regions; Coastal Zone and Wetlands by 20 world regions Supplements to FAR 1992 Climate: IS92 emissions scenarios by 7 world regions [4, 5] Impacts: Agriculture by continent (6 regions); Ocean Ecology by 3 latitude zones; Questionnaire to governments on current activities on impacts by 6 WMO regions SR: Climate Change 1994 Evaluation of IS92 emissions scenarios by 4 world regions: OECD, USSR/E. Europe, 1994 [6] China/Centrally Planned Asia and Other. Second Assessment 1995 Climate: Gridded proportional circle maps for observed climate trends (5° latitude/ longitude); Report (SAR) climate projections for 7 sub-continental regions [7, 8, 9] Impacts, Adaptations, Mitigation: Energy production statistics by 10 world regions; Forests, Wood Production and Management by three zones: Tropical, Temperate, Boreal; separate chapters by physiographic types: Deserts, Mountain Regions, Wetlands, Cryosphere, Oceans, and Coastal Zones and small islands; country case studies, Agriculture by 8 continental-scale regions; Energy supply by 8 world regions Economic and Social Dimensions: Social Costs and Response Options by 6 economic regions SR: Regional Impacts 1998 10 continental-scale regions: Africa, Arctic and Antarctic, Australasia, Europe, Latin America, [10] Middle East and Arid Asia, North America, Small Island States, Temperate Asia, Tropical Asia. Subdivisions applied in some regions; Vegetation shifts mapped by 9 biomes; Baseline (1990) Socio-Economic data provided by country and for all regions except polar. SR: Land-Use Change 1998 9 Biomes; 15 land-use categories; National and Regional case studies. and Forestry [11] SR: Aviation [12] 1999 Observed and projected emissions by 22 regional air routes; Inventories by 5 economic regions SR: Technology 2000 Country case studies; Indicators of technology transfer by 6-7 economic regions Transfer [13] SR: Emissions 2000 4 SRES world regions defined in common across integrated assessment models; 11 sub-regions; Scenarios [14] Driving Factors by 6 continental regions Third Assessment 2001 Climate: gridded observations of Climate trends; 20 example Glaciers; 9 Biomes for Carbon Cycle; Report (TAR) Circulation Regimes for model evaluation; 23 "Giorgi" regions for regional climate projections [15, 16, 17] Impacts, adaptation and vulnerability: Example projections from 32 "modified-Giorgi" regions; Basins by continent; 5 Coastal types; Urban/Rural Settlements; Insurance by economic regions; 8 continental-scale regions equivalent to 1998 Special report but with single chapter for Asia; Subdivisions used for each region (Africa, Asia and Latin America by climate zones; North America by 6 core regions and 3 border regions) Mitigation: Country examples; Developed (Annex I) and Developing (non-Annex I); Various economic regions; Policies, Measures and Instruments by 4 blocs: OECD, Economies in Transition, China and Centrally Planned Asia, and Rest of the World. SR: Ozone Layer [18] 2005 Various economic regions/countries depending on sources and uses of chemicals; SR: Carbon Capture 2005 CO2 sources by 9 economic regions; potential storage facilities: by geological formation, by oil/gas and Storage [19] wells, by ocean depth,; costs, by 4 economic groupings Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Fourth Assessment 2007 Climate: Land-use types for surface forcing of climate; Observations by 19 "Giorgi" regions; Report (AR4) Modes of variability for Model Evaluation; Attribution of climate change by 22 "Giorgi-type" [20, 21, 22] regions and by 6 ocean regions; Climate statistics for 30 "Giorgi-type" regions; PDFs of projections for 26 regions; summary graphs for 8 continental regions Impacts, adaptation and vulnerability: Studies reporting observed impacts by 7 IPCC regions; comparison of TAR and AR4 climate projections for 32 Giorgi regions; Ecosystems by 11 biomes; Agriculture by latitudinal zone; Examples of Coastal mega-Deltas; Industry and settlement by continetal region; 8 continental regions, as in TAR, but Small Islands not Small Island States; Sub- regional summary maps for each region, using physiographic, biogeographic or geographic definitions; Example vulnerability maps at sub-national scale and globally by country. Mitigation: 17 global economic regions for GDP; Energy supply by continent, by economic regions, by 3 UNFCCC groupings; Trends in CO2 emissions (and projections) , waste and carbon balance by economic regions. SR: Renewable Energy 2012 Global maps showing potential resources for renewable energy: land suitability for bioenergy Sources and Climate production, global irradiance for solar, geoethermal, hydropower, ocean waves/tidal range, wind); Change Mitigation [23] Various economic/continental regions: installed capacity (realised vs. potential), types of technologies, investment cost, cost effectiveness, various scenario-based projections; Country comparisons of deployment and uptake of technologies, share of energy market. SR: Managing the 2012 Trends in observed (tables) and projected (maps and tables) climate extremes (Tmax, Tmin, heat Risks of Extreme Events waves, heavy precipitation and dryness) by 26 sub-continental regions covering most land areas of and Disasters to the globe; Attribution studies of return periods of extreme temperatures for 15 "Giorgi-type" Advance Climate regions; Gridded global maps of projected extremes of temperature, precipitation, windspeed, dry Change Adaptation spells and soil moisture anomalies; Continental-scale estimates of projected changes in impacts of [24] extremes (floods, cyclones, coastal inundation) as well as frequencies of observed climate extremes and their estimated costs); Distinctions drawn between local, country and international/global actors with respect to risk management and its financing. Fifth Assessment 2014 Climate: Gridded global maps of observed changes in climate; Cryosphere observations from 19 Report (AR5) glacierized regions and 3 Arctic permafrost zones; Paleoclimatic reconstructions for 7 continental [25, 26, 27] regions; CO2 fluxes for 11 land and 10 ocean regions; Observed aerosol concentrations for 6 continental regions and projections for 9 regions; Detection and attribution of changes in mean and extreme climate for 7 continental and 8 ocean regions; Climate model evaluation and multi-model projections of extremes for 26 sub-continental regions; Maps and time series of seasonal and annual multi-model simulated climate changes for 19 sub-continental regions and global over 1900-2100. Impacts, adaptation and vulnerability Part A Global and sectoral aspects: Gridded global maps of water resources, species distributions, ocean productivity; Global map of 51 ocean biomes; Detection and attribution of observed impacts, key risks and vulnerabilities and adaptation synthesis by IPCC regions. Part B Regional aspects: Nine continental-scale regions, eight as in AR4 plus the ocean; Sub-regions in Africa (5), Europe (5), Asia (6), Central and South America (5 or 7); Polar (2); Small Islands (4), Oceans (7); Other disaggregation by gridded maps or countries. Mitigation: Economic statistics by development (3 or 5 categories) or by income; 5 country groupings (plus international transport) for emission-related scenario analysis (RCP5: OECD 1990 countries, Reforming Economies, Latin America and Caribbean, Middle East and Africa, Asia) with further disaggregation to 10 regions (RCP10) for regional development; Land use regions for forest (13) and agriculture (11); Most other analyses by example countries. 1. IPCC (1990c); 2. IPCC (1990a); 3. IPCC (1990b); 4. IPCC (1992b); 5. IPCC (1992a); 6. IPCC (1994) 7. IPCC (1996c); 8. IPCC (1996b); 9. IPCC (1996a); 10. IPCC (1998b); 11. IPCC (1998a); 12. IPCC (1999); 13. IPCC(2000a), 14. IPCC (2000b); 15. IPCC (2001c); 16. IPCC (2001a); 17. IPCC (2001b); 18. IPCC/TEAP (2005); 19. IPCC (2005); 20. IPCC (2007c); 21. IPCC (2007a); 22. IPCC (2007b); 23. IPCC (2012b); 24. IPCC (2012a); 25. IPCC (2014b); 26. This assessment; 27. IPCC (2014a). Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-3: Two possible entry points for thinking about vulnerability to climate change (illustrative and adapted from Füssel 2007). Context Climate change impacts perspective Vulnerability perspective Root problem Climate change Social vulnerability Policy context Climate change mitigation, Social adaptation, sustainable compensation, technical adaptation development Illustrative policy question What are the benefits of climate change How can the vulnerability of societies to mitigation? climatic hazards be reduced? Illustrative research What are the expected net impacts of Why are some groups more affected by question climate change in different regions? climatic hazards than others? Vulnerability and adaptive Adaptive capacity determines Vulnerability determines adaptive capacity vulnerability capacity Reference for adaptive Adaptation to future climate change Adaptation to current climate variability capacity Starting point of analysis Scenarios of future climate change Current vulnerability to climatic variabilty Analytical function Descriptive, positivist Explanatory, normative Main discipline Natural science Social science Meaning of vulnerability Expected net damage for a given level Susceptibility to climate change and of global climate change variability as determined by socioeconomic factors Vulnerability approach Integrated, risk-hazard Political economy Reference IPCC (2001a) Adger (1999) Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-4: Illustrative examples of adaptation experience, as well as approaches to reduce vulnerability and enhance resilience. Adaptation actions can be influenced by climate variability, extremes, and change, and by exposure and vulnerability at the scale of risk management. Many examples and case studies demonstrate complexity at the level of communities or specific regions within a country. It is at this spatial scale that complex interactions between vulnerabilities, inequalities, and climate change come to the fore. At the same time, place- based examples illustrate how larger-level drivers and stressors shape differential risks and livelihood trajectories, often mediated by institutions. Early warning systems for heat EXPOSURE AND VULNERABILITY: Factors affecting exposure and vulnerability include age, pre-existing health status, level of outdoor activity, socioeconomic factors including poverty and social isolation, access to and use of cooling, physiological and behavioral adaptation of the population, urban heat island effects, and urban infrastructure. [8.2.3-4, 11.3.3-4, 11.4.1, 11.7, 13.2.1, 19.3.2, 23.5.1, 25.3, 25.8.1, SREX Table SPM.1] CLIMATE INFORMATION AT THE GLOBAL SCALE: Observed: Very likely decrease in the number of cold days and nights and increase in the number of warm days and nights, on the global scale between 1951 and 2010. [WGI AR5 2.6.1] Medium confidence that the length and frequency of warm spells, including heat waves, has increased globally since 1950. [WGI AR5 2.6.1] Projected: Virtually certain that, in most places, there will be more hot and fewer cold temperature extremes as global mean temperatures increase, for events defined as extremes on both daily and seasonal timescales. [WGI AR5 12.4.3] CLIMATE INFORMATION AT THE REGIONAL SCALE: Observed: Likely that heat wave frequency has increased since 1950 in large parts of Europe, Asia, and Australia. [WGI AR5 2.6.1] Medium confidence in overall increase in heat waves and warm spells in North America since 1960. Insufficient evidence for assessment or spatially varying trends in heat waves or warm spells for South America and most of Africa. [SREX Table 3-2; WGI AR5 2.6.1] Projected: Likely that, by the end of the 21st century under RCP8.5 in most land regions, a current 20-year high temperature event will at least double its frequency and in many regions occur every two years or annually, while a current 20-year low temperature event will become exceedingly rare. [WGI AR5 12.4.3] Very likely more frequent and/or longer heat waves or warm spells over most land areas. [WGI AR5 12.4.3] DESCRIPTION: Heat-health early warning systems are instruments to prevent negative health impacts during heat waves. Weather forecasts are used to predict situations associated with increased mortality or morbidity. Components of effective heatwave and health warning systems include identifying weather situations that adversely affect human health, monitoring weather forecasts, communicating heatwave and prevention responses, targeting notifications to vulnerable populations, and evaluating and revising the system to increase effectiveness in a changing climate. Warning systems for heat waves have been planned and implemented broadly, for example in Europe, the United States, Asia, and Australia. [11.7.3, 24.4.6, 25.8.1, 26.6, Box 25-6] BROADER CONTEXT: Heat health warning systems can be combined with other elements of a health protection plan, for example building capacity to support communities most at risk, supporting and funding health services, and distributing public health information. In Africa, Asia, and elsewhere, early warning systems have been used to provide warning of and reduce a variety of risks, related to famine and food insecurity; flooding and other weather-related hazards; exposure to air pollution from fire; and vector- borne and food-borne disease outbreaks. [7.5.1, 11.7,15.4.2, 22.4.5, 24.4.6, 25.8.1, 26.6.3, Box 25-6] Mangrove restoration to reduce flood risks and protect shorelines from storm surge EXPOSURE AND VULNERABILITY : Loss of mangroves increases exposure of coastlines to storm surge, coastal erosion, saline intrusion, and tropical cyclones. Exposed infrastructure, livelihoods, and people are vulnerable to associated damage. Areas with development in the coastal zone, such as on small islands, can be particularly vulnerable. [5.4.3, 5.5.6, 29.7.2, Box CC-EA] CLIMATE INFORMATION AT THE GLOBAL SCALE: Observed: Likely increase in the magnitude of extreme high sea level events since 1970, mostly explained by rising mean sea Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 level. [WGI AR5 3.7.5] Low confidence in long-term (centennial) changes in tropical cyclone activity, after accounting for past changes in observing capabilities. [WGI AR5 2.6.3] Projected: Very likely significant increase in the occurrence of future sea level extremes by 2050 and 2100. [WGI AR5 13.7.2] In the 21st century, likely that the global frequency of tropical cyclones will either decrease or remain essentially unchanged. Likely increase in both global mean tropical cyclone maximum wind speed and rainfall rates. [WGI AR5 14.6] CLIMATE INFORMATION AT THE REGIONAL SCALE: Observed: Change in sea level relative to the land (relative sea level) can be significantly different from the global mean sea level change because of changes in the distribution of water in the ocean and vertical movement of the land. [WGI AR5 3.7.3] Projected: Low confidence in region-specific projections of storminess and associated storm surges. [WGI AR5 13.7.2] Projections of regional changes in sea level reach values of up to 30% above the global mean value in the Southern Ocean and around North America, and between 10% to 20% above the global mean value in equatorial regions. [WGI AR5 13.6.5] More likely than not substantial increase in the frequency of the most intense tropical cyclones in the western North Pacific and North Atlantic. [WGI AR5 14.6] DESCRIPTION: Mangrove restoration and rehabilitation has occurred in a number of locations (e.g., Vietnam, Djibouti, and Brazil) to reduce coastal flooding risks and protect shorelines from storm surge. Restored mangroves have been shown to attenuate wave height and thus reduce wave damage and erosion. They protect aquaculture industry from storm damage and reduce saltwater intrusion. [2.4.3, 5.5.4, 8.3.3, 22.4.5, 27.3.3] BROADER CONTEXT: Considered a low-regrets option benefiting sustainable development, livelihood improvement, and human well-being through improvements for food security and reduced risks from flooding, saline intrusion, wave damage, and erosion. Restoration and rehabilitation of mangroves, as well as of wetlands or deltas, is ecosystem-based adaptation that enhances ecosystem services. Synergies with mitigation given that mangrove forests represent large stores of carbon. Well-integrated ecosystem-based adaptation can be more cost effective and sustainable than non-integrated physical engineering approaches. [5.5, 8.4.2, 14.3.1, 24.6, 29.3.1, 29.7.2, 30.6.1-2, Table 5-4, Box CC-EA] Community-based adaptation and traditional practices in small island contexts EXPOSURE AND VULNERABILITY: With small land area, often low elevation coasts, and concentration of human communities and infrastructure in coastal zones, small islands are particularly vulnerable to rising sea levels and impacts such as inundation, saltwater intrusion, and shoreline change. [29.3.1, 29.3.3, 29.6.1-2, 29.7.2] CLIMATE INFORMATION AT THE GLOBAL SCALE: Observed: Likely increase in the magnitude of extreme high sea level events since 1970, mostly explained by rising mean sea level. [WGI AR5 3.7.5] Low confidence in long-term (centennial) changes in tropical cyclone activity, after accounting for past changes in observing capabilities. [WGI AR5 2.6.3] Since 1950 the number of heavy precipitation events over land has likely increased in more regions than it has decreased. [WGI AR5 2.6.2] Projected: Very likely significant increase in the occurrence of future sea level extremes by 2050 and 2100. [WGI AR5 13.7.2] In the 21st century, likely that the global frequency of tropical cyclones will either decrease or remain essentially unchanged. Likely increase in both global mean tropical cyclone maximum wind speed and rainfall rates. [WGI AR5 14.6] Globally, for short-duration precipitation events, likely shift to more intense individual storms and fewer weak storms. [WGI AR5 12.4.5] CLIMATE INFORMATION AT THE REGIONAL SCALE: Observed: Change in sea level relative to the land (relative sea level) can be significantly different from the global mean sea level change because of changes in the distribution of water in the ocean and vertical movement of the land. [WGI AR5 3.7.3] Projected: Low confidence in region-specific projections of storminess and associated storm surges. [WGI AR5 13.7.2] Projections of regional changes in sea level reach values of up to 30% above the global mean value in the Southern Ocean and around North America, and between 10% to 20% above the global mean value in equatorial regions. [WGI AR5 13.6.5] More likely than not substantial increase in the frequency of the most intense tropical cyclones in the western North Pacific and North Atlantic. [WGI AR5 14.6] DESCRIPTION: Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Traditional technologies and skills can be relevant for climate adaptation in small island contexts. In the Solomon Islands, relevant traditional practices include elevating concrete floors to keep them dry during heavy precipitation events and building low aerodynamic houses with palm leaves as roofing to avoid hazards from flying debris during cyclones, supported by perceptions that traditional construction methods are more resilient to extreme weather. In Fiji after cyclone Ami in 2003, mutual support and risk sharing formed a central pillar for community-based adaptation, with unaffected households fishing to support those with damaged homes. Participatory consultations across stakeholders and sectors within communities and capacity building taking into account traditional practices can be vital to the success of adaptation initiatives in island communities, such as in Fiji or Samoa. [29.6.2] BROADER CONTEXT: Perceptions of self-efficacy and adaptive capacity in addressing climate stress can be important in determining resilience and identifying useful solutions. The relevance of community-based adaptation principles to island communities, as a facilitating factor in adaptation planning and implementation, has been highlighted, for example with focus on empowerment and learning-by-doing, while addressing local priorities and building on local knowledge and capacity. Community-based adaptation can include measures that cut across sectors and technological, social, and institutional processes, recognizing that technology by itself is only one component of successful adaptation. [5.5.4, 29.6.2] Adaptive approaches to flood defense in Europe EXPOSURE AND VULNERABILITY : Increased exposure of persons and property in flood risk areas has contributed to increased damages from flood events over recent decades. [5.4.3-4, 5.5.5, 23.3.1, Box 5-1] CLIMATE INFORMATION AT THE GLOBAL SCALE: Observed: Likely increase in the magnitude of extreme high sea level events since 1970, mostly explained by rising mean sea level. [WGI AR5 3.7.5] Since 1950 the number of heavy precipitation events over land has likely increased in more regions than it has decreased. [WGI AR5 2.6.2] Projected: Very likely that the time-mean rate of global mean sea level rise during the 21st century will exceed the rate observed during 1971-2010 for all RCP scenarios. [WGI AR5 13.5.1] Globally, for short-duration precipitation events, likely shift to more intense individual storms and fewer weak storms. [WGI AR5 12.4.5] CLIMATE INFORMATION AT THE REGIONAL SCALE: Observed: Likely increase in the frequency or intensity of heavy precipitation in Europe, with some seasonal and/or regional variations. [WGI AR5 2.6.2] Increase in heavy precipitation in winter since the 1950s in some areas of northern Europe (medium confidence). Increase in heavy precipitation since the 1950s in some parts of west-central Europe and European Russia, especially in winter (medium confidence). [SREX Table 3-2] Increasing mean sea level with regional variations, except in the Baltic sea where the relative sea level is decreasing due to vertical crustal motion. [5.3.2, 23.2.2] Projected: Over most of the mid-latitude land-masses, extreme precipitation events will very likely be more intense and more frequent in a warmer world. [WGI AR5 12.4.5] Overall precipitation increase in northern Europe and decrease in southern Europe (medium confidence). [23.2.2] Increased extreme precipitation in northern Europe during all seasons, particularly winter, and in central Europe except in summer (high confidence). [23.2.2; SREX Table 3-3] DESCRIPTION: Several governments have made ambitious efforts to address flood risk and sea level rise over the coming century. In the Netherlands, government recommendations include soft measures preserving land from development to accommodate increased river inundation, maintaining coastal protection through beach nourishment, and ensuring necessary political- administrative, legal, and financial resources. Through a multi-stage process, the British government has also developed extensive adaptation plans to adjust and improve flood defenses in order to protect London from future storm surges and river flooding. Pathways have been analyzed for different adaptation options and decisions, depending on eventual sea level rise, with ongoing monitoring of the drivers of risk informing decisions. [5.5.4, 23.7.1, Box 5-1] BROADER CONTEXT: The Dutch plan is considered a paradigm shift, addressing coastal protection by working with nature and providing room for Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 river. The British plan incorporates iterative, adaptive decisions depending on the eventual sea level rise with numerous and diverse measures possible over the next 50-100 years to reduce risk to acceptable levels. In cities in Europe and elsewhere, the importance of strong political leadership or government champions in driving successful adaptation action has been noted. [5.5.3-4, 8.4.3, 23.7.1-2, 23.7.4, Boxes 5-1 and 26-3] Index-based insurance for agriculture in Africa EXPOSURE AND VULNERABILITY: Susceptibility to food insecurity and depletion of farmers' productive assets following crop failure. Low prevalence of insurance due to absent or poorly developed insurance markets or to amount of premium payments. The most marginalized and resource- poor especially may have limited ability to afford insurance premiums. [10.7.6, 13.3.2, Box 22-1] CLIMATE INFORMATION AT THE GLOBAL SCALE: Observed: Very likely decrease in the number of cold days and nights and increase in the number of warm days and nights, on the global scale between 1951 and 2010. [WGI AR5 2.6.1] Medium confidence that the length and frequency of warm spells, including heat waves, has increased globally since 1950. [WGI AR5 2.6.1] Since 1950 the number of heavy precipitation events over land has likely increased in more regions than it has decreased. [WGI AR5 2.6.2] Low confidence in a global-scale observed trend in drought or dryness (lack of rainfall). [WGI AR5 2.6.2] Projected: Virtually certain that, in most places, there will be more hot and fewer cold temperature extremes as global mean temperatures increase, for events defined as extremes on both daily and seasonal timescales. [WGI AR5 12.4.3] Regional to global-scale projected decreases in soil moisture and increased risk of agricultural drought are likely in presently dry regions, and are projected with medium confidence by the end of this century under the RCP8.5 scenario. [WGI AR5 12.4.5] Globally, for short-duration precipitation events, likely shift to more intense individual storms and fewer weak storms. [WGI AR5 12.4.5] CLIMATE INFORMATION AT THE REGIONAL SCALE: Observed: Medium confidence in increase in frequency of warm days and decrease in frequency of cold days and nights in southern Africa. [SREX Table 3-2] Medium confidence in increase in frequency of warm nights in northern and southern Africa. [SREX Table 3-2] Projected: Likely surface drying in southern Africa by the end of this century under RCP8.5 (high confidence). [WGI AR5 12.4.5] Likely increase in warm days and nights and decrease in cold days and nights in all regions of Africa (high confidence). Increase in warm days largest in summer and fall (medium confidence). [Table SREX 3-3] Likely more frequent and/or longer heat waves and warm spells in Africa (high confidence). [Table SREX 3-3] DESCRIPTION: A recently introduced mechanism that has been piloted in a number of rural locations, including in Malawi, Sudan, and Ethiopia, as well as in India. When physical conditions reach a particular predetermined threshold where significant losses are expected to occur--weather conditions such as excessively high or low cumulative rainfall or temperature peaks--the insurance pays out. [9.4.2, 13.3.2, 15.4.4, Box 22-1] BROADER CONTEXT: Index-based weather insurance is considered well-suited to the agricultural sector in developing countries. The mechanism allows risk to be shared across communities, with costs spread over time, while overcoming obstacles to traditional agricultural and disaster insurance markets. It can be integrated with other strategies such as micro-finance and social protection programs. Risk-based premiums can help encourage adaptive responses and foster risk awareness and risk reduction by providing financial incentives to policyholders to reduce their risk profile. Challenges can be associated with limited availability of accurate weather data and difficulties in establishing which weather conditions cause losses. Basis risk (i.e., farmers suffer losses but no payout is triggered based on weather data) can promote distrust. There can also be difficulty in scaling up pilot schemes. Insurance for work programs can enable cash-poor farmers to work for insurance premiums by engaging in community- identified disaster risk reduction projects. [10.7.4-6, 13.3.2, 15.4.4, Table 10-7, Box 22-1, Box 25-7] Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Relocation of agricultural industries in Australia EXPOSURE AND VULNERABILITY : Crops sensitive to changing patterns of temperature, rainfall, and water availability. [7.3, 7.5.2] CLIMATE INFORMATION AT THE GLOBAL SCALE: Observed: Very likely decrease in the number of cold days and nights and increase in the number of warm days and nights, on the global scale between 1951 and 2010. [WGI AR5 2.6.1] Medium confidence that the length and frequency of warm spells, including heat waves, has increased globally since 1950. [WGI AR5 2.6.1] Medium confidence in precipitation change over global land areas since 1950. [WGI AR5 2.5.1] Since 1950 the number of heavy precipitation events over land has likely increased in more regions than it has decreased. [WGI AR5 2.6.2] Low confidence in a global-scale observed trend in drought or dryness (lack of rainfall). [WGI AR5 2.6.2] Projected: Virtually certain that, in most places, there will be more hot and fewer cold temperature extremes as global mean temperatures increase, for events defined as extremes on both daily and seasonal timescales. [WGI AR5 12.4.3] Virtually certain increase in global precipitation as global mean surface temperature increases. [WGI AR5 12.4.1] Regional to global-scale projected decreases in soil moisture and increased risk of agricultural drought are likely in presently dry regions, and are projected with medium confidence by the end of this century under the RCP8.5 scenario. [WGI AR5 12.4.5] Globally, for short-duration precipitation events, likely shift to more intense individual storms and fewer weak storms. [WGI AR5 12.4.5] CLIMATE INFORMATION AT THE REGIONAL SCALE: Observed: Cool extremes rarer and hot extremes more frequent and intense over Australia and New Zealand, since 1950 (high confidence). [Table 25-1] Likely increase in heat wave frequency since 1950 in large parts of Australia. [WGI AR5 2.6.1] Late autumn/winter decreases in precipitation in Southwestern Australia since the 1970s and Southeastern Australia since the mid-1990s, and annual increases in precipitation in Northwestern Australia since the 1950s (very high confidence). [Table 25-1] Mixed or insignificant trends in annual daily precipitation extremes, but a tendency to significant increase in annual intensity of heavy precipitation in recent decades for sub-daily events in Australia (high confidence). [Table 25-1] Projected: Hot days and nights more frequent and cold days and nights less frequent during the 21st century in Australia and New Zealand (high confidence). [Table 25-1] Annual decline in precipitation over southwestern Australia (high confidence) and elsewhere in southern Australia (medium confidence). Reductions strongest in the winter half-year (high confidence). [Table 25-1] Increase in most regions in the intensity of rare daily rainfall extremes and in sub-daily extremes (medium confidence) in Australia and New Zealand. [Table 25-1] Drought occurrence to increase in Southern Australia (medium confidence). [Table 25-1] Snow depth and snow area to decline in Australia (very high confidence). [Table 25-1] Freshwater resources projected to decline in far southeastern and far southwest Australia (high confidence). [25.5.2] DESCRIPTION: Industries and individual farmers are relocating parts of their operations, for example for rice, wine, or peanuts in Australia, or are changing land use in situ in response to recent climate change or expectations of future change. For example, there has been some switching from grazing to cropping in South Australia. Adaptive movement of crops has also occurred elsewhere. [7.5.1, 25.7.2, Table 9-7, Box 25-5] BROADER CONTEXT: Considered transformational adaptation in response to impacts of climate change. Positive or negative implications for the wider communities in origin and destination regions. [25.7.2, Box 25-5] Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-5: Dimensions of assessments of impacts and vulnerability and of adaptation drawn upon to serve different target fields (cf. Table 21-1). Scales refer to the level of aggregation at which study results are presented. Entries are illustrations of different types of study approaches reported and evaluated in this volume, with references given both to the original studies and to the chapters in which they are cited. Aspects of some of the studies in this table are also alluded to in Section 21.5. Approach/field: Impacts/vulnerability Adaptation Target field Scale: Resource availability1,2,3 Adaptation costs4,5,6,7,12 -­ Policy negotiations Impact costs4,5,6,7 -­ Development aid Global Vulnerability/risk mapping8,9,10 -­ Disaster planning Hotspots analysis11 -­ Capacity building Observed impacts13,14,15 Adaptation costs5 -­ Capacity building Continental/ Future biophysical impacts16,17 Modelled adaptation19 -­ International law biome Impact costs5,16 -­ Policy negotiations Vulnerability/risk mapping18 -­ Regional development Observed impacts20,21,22 Observed adaptation26 -­ National adaptation Future impacts/risks23,24 Adaptation assessment24,27 plan/strategy National/ Vulnerability assessment24 -­ Nat. Communication state/province Impact costs25 -­ Legal requirement -­ Regulation Hazard/risk mapping28 Adaptation cost28 -­ Spatial planning Municipality/ Pest/disease risk mapping29 Urban adaptation30,31 -­ Extension services basin/patch/ Urban risks/vulnerabilities30 -­ Water utilities delta/farm -­ Private sector Site/field/tree/ Field experiments32 Coping studies33,34 -­ Individual actors floodplain/ Economic modelling35 -­ Local planners household Agent-based modelling36 Notes for Table 21-5 1 Global terrestrial water balance, in the Water Model Intercomparison Project (Haddeland et al., 2011), see chapter 3 2 Global dynamic vegetation model intercomparison (Sitch et al., 2008), see chapter 4 3 Impacts on agriculture, coasts, water resources, ecosystems and health in the Inter-Sectoral Impact Model Intercomparison Project (ISI-MIP Schiermeier, 2012), see chapter 19 4 UNFCCC study to estimate the aggregate cost of adaptation (UNFCCC, 2007), which is critiqued by Parry (2009) and Fankhauser (2010) 5 The Economics of Adaptation to Climate Change study (World Bank, 2010) 6 Chapter 17 provides a thorough evaluation of global modelling studies (see also chapters 14 and 16) 7 Impacts on agriculture and costs of adaptation (e.g. Nelson et al., 2009), see chapter 7 8 Can we avoid dangerous climate change? (AVOID) programme and Quantifying and Understanding the Earth System (QUEST) Global-scale impacts of climate change (GSI) project (Arnell et al., 2013), see chapter 19 9 OECD project on Cities and Climate Change (Hanson et al., 2011), see chapters 5, 23, 24 and 26 10 For critical reviews of global vulnerability studies, see Füssel (2010) and Preston et al.(2011) 11 A discussion of hotspots can be found in section 21.5.1.2 12 Adaptation costs for climate change-related human health impacts (Ebi, 2008), see chapter 17 13 Satellite monitoring of sea ice over polar regions (Comiso and Nishio, 2008), see also Vaughan et al. (2014) and chapters 18 and 28 14 Satellite monitoring of vegetation growth (e.g., Piao et al., 2011) and phenology (e.g., Heumann et al., 2007), see chapters 4 and 18 15 Meta-analysis of range shifts in terrestrial organisms (e.g., Chen et al., 2011), see chapters 4 and 18 16 Physical and economic impacts of future climate change in Europe (Ciscar et al., 2011), see chapter 23 17 Impacts on crop yields in West Africa (Roudier et al., 2011), see chapter 22 18 Climate change integrated methodology for cross-sectoral adaptation and vulnerability in Europe (CLIMSAVE) project (Harrison et al., 2012), see chapter 23 19 Modelling agricultural management under climate change in sub-Saharan Africa (Waha et al., 2013) 20 Satellite monitoring of lake levels in China (Wang et al., 2013) Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 21 Satellite monitoring of rice phenology in India (Singh et al., 2006), see chapter 18 22 UK Climate Change Risk Assessment (CCRA, 2012), see chapter 23 23 United States Global Change Research Program second (Karl et al., 2009) and third (in review) national climate change impact assessments, see chapter 26 24 The Global Environment Facility (GEF)-funded Assessments of Impacts and Adaptations to Climate Change (AIACC) program addressed impacts and vulnerability (Leary et al., 2008b) and adaptation (Leary et al., 2008a) in developing countries, see chapter 27 25 Economics of Climate Change national studies in Kenya and Tanzania (SEI, 2009; GCAP, 2011), see chapter 22 26 Sowing dates of various crops in Finland (Kaukoranta and Hakala, 2008), and see chapter 18 27 Finnish Climate Change Adaptation Research Programme (ISTO) Synthesis Report (Ruuhela, 2012) 28 Urban flood risk and adaptation cost, Finland (Perrels et al., 2010) 29 See Garrett (2013) for a specific example of a risk analysis, or Sutherst (2011) for a review and see chapter 25 30 New York City coastal adaptation (Rosenzweig et al., 2011), see chapters 8 and 26 31 Bangkok Assessment Report of Climate Change (BMA/GLF/UNEP, 2009), see chapters 8 and 24 32 Field, chamber and laboratory plant response experiments (e.g., Long et al., 2006; Hyvönen et al., 2007; Wittig et al., 2009; Craufurd et al., 2013), see chapters 4 and 7 33 Farming response to irrigation water scarcity in China (Liu et al., 2008) and see chapter 13 34 Farmers' mechanisms for coping with hurricanes in Jamaica (Campbell and Beckford, 2009) and see chapter 29 35 Modelling micro-insurance of subsistence farmers for drought losses in Ethiopia (Meze-Hausken et al., 2009), see chapter 14 36 Simulating adaptive behaviour of farming communities in the Philippines (Acosta-Michlik and Espaldon, 2008), see chapter 24 Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-6: Reliability of climate information on temperature and precipitation over a range of spatial and temporal scales. Reliability is assigned to one of seven broad categories from Very High (VH) to Medium (M) through to Very Low (VL). Scale Temporal Annual Seasonal-Monthly Daily Spatial Variable Temp Precip Temp Precip Temp Precip Era Past VH H VH H N/A N/A Global Future VH direction H direction VH direction H direction N/A N/A change H amount MH amount H amount MH amount VH-H VH-H depends H-L depends on VH-H depends H-L depends on H-L depends depends on Past on observation observation on observation observation on observation Regional, observation availability availability availability availability availability Large availability river basin H-L depends on H-L depends on H-L depends Future VH direction VH direction VH direction capture of capture of on capture of change H amount MH amount MH amount processes processes processes VH-H VH-H depends H-L depends on VH-H depends H-L depends on H-VL depends depends on Past on observation observation on observation observation on observation observation National, availability availability availability availability availability availability State H-L depends on H-L depends on H-VL depends Future VH direction VH direction H direction capture of capture of on capture of change MH amount MH amount MH amount processes processes processes VH-M depends H-VL depends VH-M depends H-VL depends H-ML depends H-VL depends Past on observation on observation on observation on observation on observation on observation City, availability availability availability availability availability availability County H-VL depends H-VL depends M-VL depends Future H direction H direction H direction on capture of on capture of on capture of change MH amount MH amount M amount processes processes processes VH-ML depends H-VL depends VH-ML depends H-VL depends H-ML depends H-VL depends Past on observation on observation on observation on observation on observation on observation Village, availability availability availability availability availability availability Site/field H-VL depends H-VL depends M-VL depends Future H direction H direction H direction on capture of on capture of on capture of change MH amount MH amount M amount processes processes processes " Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-7: An assessment of observed and projected future changes in temperature and precipitation extremes over 26 sub-continental regions as defined in the SREX report (IPCC 2012; see also Figure 21.4 and Table SM21.2). Confidence levels are indicated by colour coding of the symbols. Likelihood terms are given only for high confidence statements and are specified in the text. Observed trends in temperature and precipitation extremes, including dryness, are generally calculated from 1950, using the period 1961-1990 as a baseline (see Box 3.1 chapter 3 of IPCC (2012a)). The future changes are derived from global and regional climate model projections of the climate of 2071-2100 compared with 1961-1990 or 2080-2100 compared with 1980-2000. Table entries are summaries of information in Tables 3.2 and 3.3 of IPCC (2012a) supplemented with or superseded by material from Chapters 2 (section 2.6 and Table 2.13) and 14 (section 14.4) of the IPCC AR5 WG1 report and Table 25-1 of the IPCC WG2 report. The source(s) of information for each entry are indicated by the superscripts a (Table 3.2 of IPCC, 2012a), b (Table 3.3 of IPCC, 2012a), c (Chapter 2 (section 2.6 and Table 2.13) IPCC AR5 WG1 report), d (Chapter 14 (section 14.4) of the IPCC AR5 WG1 report) and e (Table 25-1 of the IPCC WG2 report). Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 21-8: Leading knowledge gaps and related research needs. Knowledge Gap Research need 1 There is no clear understanding of how to integrate the Research is needed to distinguish the relative stochastic and diversity of climate change projections data. The full deterministic sources of variability and change as a function associated uncertainty is weakly characterized and of scale, variable, and application. The need is to develop quantifying how much of an observed or simulated climate further and build on physical understanding of the drivers of change is due to internal variability or external forcings is climate variability and change and how to represent these difficult in many situations. Collectively this results in data realistically within models to understand the source of the products with differing time and space resolution, differing spread and any contradictions in the regional projections at dependencies and assumptions and that can have conflicting scales relevant to users and then to provide guidance on a messages. At present individual products are plausible and likely range of outcomes within which the true response mostly defensible in-so-far as they have physical basis would be expected to lie. Similarly, there is a need is to within the assumptions of the method. However, at articulate the real inherent uncertainty within climate decision-relevant scales understanding where (or whether) projection data and to understand when climate information the true outcome will lie within the range of the products is useful at the scales of need. This also requires stronger collectively is often not possible and thus the products are dialogues with users of climate information to inform often not strongly actionable. choices of variables and ways to characterize envelopes of risk and uncertainties. 2. The growth of multi-model, multi-method, and multi- Methodological and conceptual advances are needed to generational data for climate projections creates confusion facilitate the synthesis of diverse data sets on different for the IAV community. The lack of a clear approach to scales from methods with different assumptions, and to handling this diversity leads to choosing one or other subset, integrate these into cohesive and defensible understanding and where one choice may substantially alter the IAV of projected regional change. conclusion compared to a different subset. 3 The attributes of regional climate change through which The research need is to be able to demonstrate how to impacts are manifest, such as the intensity, persistence, unpack the regional projections into terms relevant for distribution, recurrence, and frequency of weather events is impacts and adaptation. For example, how is the shape of poorly understood. The information conveyed to the the distribution of weather events changing (not just the adaptation community is dominated by aggregates in time extremes), or how stable are the critical global and space (e.g. SREX regional averages, or time averages), teleconnection patterns that contribute to the variability of a which hide the important attributes underlying these region? aggregated changes. In part this is a consequence of (1) above. 4 The historical record for many regions, especially those The research need is to integrate the multiplicity of regions most vulnerable to climate change, is poor to the historical data as represented by the raw observations extent that the historical record is at best an estimate with processed gridded products (e.g. CRU and GPCP), satellite unknown uncertainty. This severely undermines the data, and reanalysis data sets. Involving national scientists development of regional change analysis, limits the with their inherent local knowledge and rescue and evaluation of model skill, and presents a weak baseline digitization of the many national archives still inaccessible against which to assess change signals or to develop to the wider research community would significantly impacts, adaptation or vulnerability baselines. enhance this research activity. 5 Impact model sensitivity studies and inter-comparison Intensified efforts are needed to refine, test and inter- exercises are beginning to reveal fundamental flaws and compare impact models over a wider range of sectors and omissions in some impact models in the representation of environments than hitherto. These should be supported, key processes that are expected to be important under where applicable, by targeted field, chamber and laboratory projected climate changes. For example, high temperature experiments under controlled atmospheric composition and constraints, and CO2 and drought effects on agricultural climate conditions, to improve understanding of key yields are poorly represented in many crop models. physical, biological and chemical processes operating in changed environments. Such experiments are needed across a range of terrestrial and aquatic biogeographical zones in different regions of the world Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 6 New global scenarios are under development, based on Extended SSPs for major sub-continental regions of the climate projections for different representative world, including variables that define aspects of adaptive concentration pathways (RCPs) and socio-economic capacity and guidance on how to combine RCP-based scenarios based on shared socio-economic pathways (SSPs). regional climate projections with regional SSPs and SPAs to However, there is currently little or no guidance on how form plausible regional scenarios for application in IAV these projections are to be accessed or applied in IAV analysis. studies. Moreover, as yet, quantitative SSPs are available only for large regions (basic SSPs), and regional SSPs that are consistent with the global SSPs (extended SSPs) along with scenarios that include mitigation and adaptation policies (shared policy assumptions SPAs) have not yet been developed., 7 The determinants and regional variability of vulnerability, Case studies and underlying theory of these features of exposure and adaptive capacity are not well understood, and societies, and documentation of the effectiveness of actions methods for projecting changes in them are under- taken are needed in conjunction with methods development developed. Furthermore, given these lacks of understanding for projections. More attention placed on determining their uncertainties of these three elements are poorly uncertainties, in national and regional assessments. characterized and quantified Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-1: Specification of the world regions described in chapters 22-30 of this volume. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-2: Observed and projected changes in annual average temperature and precipitation. (Top panel, left) Observed temperature trends from 1901-2012 determined by linear regression. [WGI AR5 Figures SPM.1 and 2.21] (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression. [WGI AR5 Figure SPM.2] For observed temperature and precipitation, trends have been calculated where sufficient data permits a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level. Diagonal lines indicate areas where change is not significant. (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and >90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in >66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See alsoAnnex I of WGI AR5. [Boxes 21-3 and CC-RC] Subject to Final Copyedit 85 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-3: Regional average change in seasonal and annual mean temperature and precipitation over five sub- regions covering South and Central America for the period 2071-2100 relative to 1961-90 in GCM projections from 35 CMIP5 ensemble under four RCP scenarios (van Vuuren et al., 2011) compared with GCM projections from 22 CMIP3 ensemble under three SRES scenarios (IPCC, 2000); see Table 21-1 for details of the relationship between the SRES and RCP scenarios. Regional averages are based on SREX region definitions (see Figure 21-3). Temperature changes are given in C and precipitation changes in mm/day with axes scaled relative to the maximum changes projected across the range of models. The models which generated the data displayed are listed in supplementary material Table SM21-3. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 86 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-4: CMIP5 ensemble median ratio of local:global average temperature change in the period 2071-2100 relative to 1961-90 under the RCP8.5 emissions/concentrations scenario. The values are displayed on a common 2.5x3.75 grid onto which each models data were regridded and they were calculated as follows: 1) for each model the local change was calculated between 1961 and 1990 at each grid cell, and is divided by the global average change in that model projection over the same period; 2) the median ratio value across all models at each grid cell is identified and shown. Data used are from the 35 CMIP5 models for which monthly projections were available under RCP8.5 which are listed in supplementary Table 21-3. Overplotted polygons indicate the SREX regions (IPCC, 2012) used to define the sub-regions used to summarise information in Chapters 21 and some of the subsequent regional chapters. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-5: Evolution of the 5%, 17%, 33%, 50%, 66%, 83% and 95% percentiles of the distribution functions for annual surface air temperature changes (upper panel) and annual precipitation (lower panel) for the Giorgi-Francisco (2000) regions and the globe with the SRES A1B forcing scenario combining results from a perturbed physics ensemble and the CMIP3 ensemble. Twenty year means relative to the 1961-1990 baseline are plotted in decadal steps using a common y-axis scale. The 5%, 50% and 95% percentile values for the period 2080-2099 are displayed for each region. (From Harris et al. 2012). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-6: Linear changes (i.e. changes obtained by fitting the time series at each grid point with straight lines) of annual precipitation during the 2001-2050 period from 10 individual RCM experiments and the MME mean under the A1B emission scenario. The top middle panels also account for projected land cover changes (see Paeth et al. 2011 for further explanation). Note that the REMO trends in both panels arise from a three-member ensemble whereas all other RCMs are represented by one single simulation. Trends statistically significant at the 95% level are marked by black dots. From Paeth et al. (2011). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 89 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-7: The frequency of 'warm days' (defined here as the 90th percentile daily maximum temperature during a baseline period of 1961-1990) projected for the 2071-2100 period by 26 CMIP5 GCMs for North America. Map: Ensemble median frequency of 'warm days' during 2071-2100 under RCP8.5. Graphs: Box-and-whisker plots indicate the range of regionally-averaged 'hot-day' frequency by 2041-2070 and 2071-2100 under RCPs 4.5 and 8.5 across the 26 CMIP5 models for each SREX sub-regions in North America. Boxes represent inter-quartile range and whiskers indicate full range of projections across the ensemble. The baseline frequency of warm days of 10% is represented on the graphs by the dashed line. A full list of CMIP5 models for which data is shown here can be found in supplementary material Table SM21-4. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 90 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-8: The frequency of 'very wet days' (defined here as the 90th percentile of daily precipitation on wet days during a baseline period of 1961-1990 with wet days defined as days with 1mm of precipitation or more) projected for the 2071-2100 period by 26 CMIP5 GCMs for Asia. Map: Ensemble median frequency of 'very wet days' during 2071-2100 under RCP8.5. Graphs: Box-and-whisker plots indicate the range of regionally-averaged 'very wet day' frequency by 2041-2070 and 2071-2100 under RCPs 4.5 and 8.5 across the 26 CMIP5 models for each SREX sub- regions in Asia Boxes represent inter-quartile range and whiskers indicate full range of projections across the ensemble. The baseline frequency of Very wet days of 10% is represented on the graphs by the dashed line. A full list of CMIP5 models for which data is shown here can be found in supplementary material Table SM21-4. (Note the WMO Expert Team on Climate Change Detection Indices defines very wet days threshold as the 95%-ile daily precipitation event. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 91 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-9: Mean (top panels) and standard deviation (bottom panels) in future-minus-present (2050s minus 1990s) MDA8 summer ozone concentrations across (left-hand panels) all seven experiments (five regional and 2 global) and for comparison purposes (right hand panels), not including the WSU experiment (which simulated July only conditions). The different experiments use different pollutant emission and SRES GHG emission scenarios. The pollutant emissions are the same in the present and future simulations (from Weaver et al., 2009). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 92 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-10: Growth rates from 1990-2008 of international trade, its embodied CO2 emissions and net emissions transfers from Annex B and non-Annex B countries compared to other global macrovariables, all indexed to 1990 (Peters et al., 2011). Annex B and non-Annex B Parties to the UNFCCC are listed in the supplementary material. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 93 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-11: Central map: Marine exclusive environmental zones (EEZs dashed lines) of Canada, Greenland/Denmark, Norway, Russia, and the USA, and location of the Northwest Passage, Northern Sea Route, Trans-Polar Route, and international high seas within the IMO Guidelines Boundary for Arctic shipping (thick black border). After Stephenson et al. (2013). Peripheral monthly maps: Projected change in accessibility of maritime and land-based transportation by mid-century (2045-2059 relative to 2000-2014) using the Arctic Transport Accessibility Model and CCSM3 climate and sea ice estimates assuming an SRES A1B scenario. Dark blue areas denote new maritime access to Polar Class 6 vessels (light icebreaker); white areas remain inaccessible. Red delimits areas of lost winter road potential for ground vehicles exceeding 2 metric tonnes (Stephenson et al., 2011). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 94 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 21-12: Time series of seasonally averaged climate indices representing three modes of large-scale climate variability: (a) the tropical September to January Pacific Walker Circulation (PWC); (b) the December to March North Atlantic Oscillation (NAO); (c) the December to March Pacific North America (PNA) pattern. Indices (as defined in Brönnimann etal. (2009) are calculated (with respect to the overlapping 1989 1999 period) from various observed, reanalysis and model sources: statistical reconstructions of the PWC, the PNA and the NAO, see Brönnimann et al. (2009) for details, (all cyan); 20CR (pink); NCEP NCAR reanalyses (NNR; dark blue); ERA-40 (green); ERA-Interim (orange). The black line and grey shading represent the ensemble mean and spread from a climate model ensemble with a lower boundary condition of observed seas-surface temperatures and sea-ice from the HadISST dataset (Rayner et al. 2003), see Brönnimann etal. (2009) for details. The model results provide a measure of the predictability of these modes of variability from sea-surface temperature and sea-ice alone and demonstrate that the reanalyses have significantly higher skill in reproduces these modes of variability. [Illustration to be redrawn to conform to IPCC publication specifications.] Figure RC-1: Observed and projected changes in global annual average temperature. Values are expressed relative to 1986-2005. Black lines show the GISTEMP, NCDC-MLOST, and HadCRUT4.2 estimates from observational measurements. Colored shading denotes the +/-1.64 standard deviation range based on simulations from 32 models for RCP2.6 (blue) and 39 models for RCP8.5 (red). Blue and red lines denote the scenario mean for RCP2.6 and RCP8.5, respectively. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 95 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure RC-2: Observed and projected changes in annual average temperature. (A) Observed temperature trends from 1901-2012 are determined by linear regression. Trends have been calculated where sufficient data permit a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level (after accounting for autocorrelation effects on significance testing). Diagonal lines indicate areas where change is not significant. Observed data are from WGI AR5 Figures SPM.1 and 2.21. The range of grid-point values is -0.53 to +2.50°C over period. (B) CMIP5 multi-model mean projections of annual average temperature changes for 2046-2065 and 2081-2100 under RCP2.6 and RCP8.5, relative to 1986-2005. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability and >90% of models agree on the sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on the sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on the sign of change. Colors with diagonal lines indicate areas with little or no change, where >66% of models show change less than the baseline variability (although there may be significant change at shorter timescales such as seasons, months, or days). Analysis uses model data from WGI AR5 Figure SPM.8, Box 12.1, and Annex I. The range of grid-point values for the multi-model mean is: +0.19 to +4.08C for mid-21st century of RCP2.6; +0.06 to +3.85C for late-21st century of RCP2.6; +0.70 to +7.04C for mid-21st century of RCP8.5; and +1.38 to +11.71°C for late-21st century of RCP8.5. Subject to Final Copyedit 96 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 21 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure RC-3: Observed and projected changes in annual average precipitation. (A) Observed precipitation trends from 1951-2010 are determined by linear regression. Trends have been calculated where sufficient data permit a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level (after accounting for autocorrelation effects on significance testing). Diagonal lines indicate areas where change is not significant. Observed data are from WGI AR5 Figures SPM.2. The range of grid-point values is -185 to +111 mm/year/decade. (B) CMIP5 multi-model mean projections of annual average precipitation changes for 2046-2065 and 2081-2100 under RCP2.6 and RCP8.5, relative to 1986-2005. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability and >90% of models agree on the sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on the sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on the sign of change. Colors with diagonal lines indicate areas with little or no change, where >66% of models show change less than the baseline variability (although there may be significant change at shorter timescales such as seasons, months, or days). Analysis uses model data from WGI AR5 Figure SPM.8, Box 12.1, and Annex I. The range of grid-point values for the multi-model mean is: -10 to +24% for mid- 21st century of RCP2.6; -9 to +22% for late-21st century of RCP2.6; -19 to +57% for mid-21st century of RCP8.5; and -34 to +112% for late-21st century of RCP8.5. Subject to Final Copyedit 97 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 22. Africa Coordinating Lead Authors Isabelle Niang (Senegal), Oliver C. Ruppel (Namibia) Lead Authors Mohamed Abdrabo (Egypt), Ama Essel (Ghana), Christopher Lennard (South Africa), Jonathan Padgham (USA), Penny Urquhart (South Africa) Contributing Authors Ibidun Adelekan (Nigeria), Sally Archibald (South Africa), Armineh Barkhordarian (Germany), Jane Battersby (South Africa), Michael Balinga (Cameroon), Eren Bilir (USA), Marshall Burke (USA), Mohammed Chahed (Tunisia), Monalisa Chatterjee (USA / India), Chineke Theo Chidiezie (Nigeria), Katrien Descheemaeker (Netherlands), Houria Djoudi (Algeria), Kristie L. Ebi (USA), Papa Demba Fall (Senegal), Ricardo Fuentes (Mexico), Rebecca Garland (South Africa), Fatou Gaye (The Gambia), Karim Hilmi (Morocco), Emiloa Gbobaniyi (Nigeria), Patrick Gonzalez (USA), Blane Harvey (UK), Mary Hayden (USA), Andreas Hemp (Germany), Guy Jobbins (UK), Jennifer Johnson (USA), David Lobell (USA), Bruno Locatelli (France), Eva Ludi (UK), Lars Otto Naess (UK), Mzime R. Ndebele-Murisa (Zimbabwe), Aminata Ndiaye (Senegal), Andrew Newsham (UK), Sirra Njai (The Gambia), Johnson Nkem (Cameroon), Jane Mukarugwiza Olwoch (South Africa), Pieter Pauw (Netherlands), Emilia Pramova (Bulgaria), Marie-Louise Rakotondrafara (Madagascar), Clionadh Raleigh (Ireland), Debra Roberts (South Africa), Carla Roncoli (USA), Aissa Toure Sarr (Senegal), Michael Henry Schleyer (South Africa), Lena Schulte-Uebbing (Germany), Roland Schulze (South Africa), Hussen Seid (Ethiopia), Sheona Shackleton (South Africa), Mxolisi Shongwe (Swaziland), Dáithí Stone (Canada / South Africa / USA), David Thomas (UK), Okoro Ugochukwu (Nigeria), Dike Victor (Nigeria), Katharine Vincent (South Africa), Koko Warner (Germany), Sidat Yaffa (The Gambia) Review Editors Pauline Dube (Botswana), Neil Leary (USA) Volunteer Chapter Scientist Lena Schulte-Uebbing (Germany) Contents Executive Summary 22.1. Introduction 22.1.1. Structure of the Regions 22.1.2. Major Conclusions from Previous Assessments 22.1.2.1. Regional Special Report and Assessment Reports 22.1.2.2. Special Report on Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX) 22.2. Observed Climate Trends and Future Projections 22.2.1. Temperature 22.2.1.1. Observed Trends 22.2.1.2. Projected Trends 22.2.2. Precipitation 22.2.2.1. Observed Changes 22.2.2.2. Projected Changes 22.2.3. Observed and Projected Changes in Extreme Temperature and Rainfall 22.3. Vulnerability and Impacts Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.3.1. Socioeconomic and Environmental Context Influencing Vulnerability and Adaptive Capacity 22.3.2. Ecosystems 22.3.2.1. Terrestrial Ecosystems 22.3.2.2. Freshwater Ecosystems 22.3.2.3. Coastal and Ocean Systems 22.3.3. Water Resources 22.3.4. Agriculture and Food Security 22.3.4.1. Crops 22.3.4.2. Livestock 22.3.4.3. Agricultural Pests, Diseases, and Weeds 22.3.4.4. Fisheries 22.3.4.5. Food Security 22.3.5. Health 22.3.5.1. Introduction 22.3.5.2. Food- and Water-Borne Diseases 22.3.5.3. Nutrition 22.3.5.4. Vector-Borne Diseases and Other Climate-Sensitive Health Outcomes 22.3.6. Urbanization 22.4. Adaptation 22.4.1. Introduction 22.4.2. Adaptation Needs, Gaps, and Adaptive Capacity 22.4.3. Adaptation, Equity, and Sustainable Development 22.4.4. Experiences in Building the Governance System for Adaptation, and Lessons Learned 22.4.4.1. Introduction 22.4.4.2. Regional and National Adaptation Planning and Implementation 22.4.4.3. Institutional Frameworks for Adaptation 22.4.4.4. Sub-National Adaptation Governance 22.4.4.5. Community-Based Adaptation and Local Institutions 22.4.4.6. Adaptation Decisionmaking and Monitoring 22.4.5. Experiences with Adaptation Measures in Africa and Lessons Learned 22.4.5.1. Overview 22.4.5.2. Climate Risk Reduction, Risk Transfer, and Livelihood Diversification 22.4.5.3. Adaptation as a Participatory Learning Process 22.4.5.4. Knowledge Development and Sharing 22.4.5.5. Communication, Education, and Capacity Development 22.4.5.6. Ecosystem Services, Biodiversity, and Natural Resource Management 22.4.5.7. Technological and Infrastructural Adaptation Responses 22.4.5.8. Maladaptation Risks 22.4.6. Barriers and Limits to Adaptation in Africa 22.5. Key Risks for Africa 22.6. Emerging Issues 22.6.1. Human Security 22.6.1.1 Violent Conflict 22.6.1.2 Migration 22.6.2. Integrated Adaptation / Mitigation Approaches 22.6.3. Biofuels and Land Use 22.6.4. Climate Finance and Management 22.7. Research Gaps References Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter Boxes 22-1. Experience with Index-Based Weather Insurance in Africa 22-2. African Success Story: Integrating Trees into Annual Cropping Systems Frequently Asked Questions 22.1: How could climate change impact food security in Africa? 22.2: What role does climate change play with regard to violent conflict in Africa? Executive Summary Evidence of warming over land regions across Africa, consistent with anthropogenic climate change, has increased (high confidence). Decadal analyses of temperatures strongly point to an increased warming trend across the continent over the last 50-100 years. [22.2.1.1] Mean annual temperature rise over Africa, relative to the late 20th Century mean annual temperature, is likely to exceed 2° C in the A1B and A2 scenarios by the end of this century (medium confidence). Warming projections under medium scenarios indicate that extensive areas of Africa will exceed 2° C by the last two decades of this century relative to the late 20th Century mean annual temperature and all of Africa under high emission scenarios. Under a high RCP, that exceedence could occur by mid-century across much of Africa and reach between 3 and 6° C by the end of the century. It is likely that land temperatures over Africa will rise faster than the global land average, particularly in the more arid regions, and that the rate of increase in minimum temperatures will exceed that of maximum temperatures. [22.2.1.2] A reduction in precipitation is likely over Northern Africa and the south-western parts of South Africa by the end of the 21st Century under the A1B and A2 scenarios (medium to high confidence). Projected rainfall change over sub-Sarahan Africa in the mid- and late 21st Century is uncertain. In regions of high or complex topography such as the Ethopian Highlands, downscaled projections indicate likely increases in rainfall and extreme rainfall by the end of the 21st Century. [22.2.2.2, 22.2.3] African ecosystems are already being affected by climate change, and futre impacts are expected to be substantial (high confidence). There is emerging evidence on shifting ranges of some species and ecosystems due to elevated CO2 and climate change, beyond the effects of land-use change and other non-climate stressors (high confidence). Ocean ecosystems, in particular coral reefs, will be affected by ocean acidification and warming as well as changes in ocean upwellings, thus negatively affecting economic sectors such as fisheries (medium confidence). [22.3.2, Table 22-3] Climate change will amplify existing stress on water availability in Africa (high confidence). Water resources are subjected to high hydro-climatic variability over space and time, and are a key constraint on the continent s continued economic development. The impacts of climate change will be superimposed onto already water-stressed catchments with complex land uses, engineered water systems, and a strong historical socio-political and economic footprint. Strategies that integrate land and water management, and disaster risk reduction, within a framework of emerging climate change risks would bolster resilient development in the face of projected impacts of climate change. [22.3.2.2, 22.3.3] Climate change will interact with non-climate drivers and stressors to exacerbate vulnerability of agricultural systems, particularly in semi-arid areas (high confidence). Increasing temperatures and changes in precipitation are very likely to reduce cereal crop productivity. This will have strong adverse effects on food security. New evidence is also emerging that high-value perennial crops could also be adversely affected by temperature rise (medium confidence). Pest, weed and disease pressure on crops and livestock is expected to increase as a result of climate change combined with other factors (low confidence). Moreover, new challenges to food security are emerging as a result of strong urbanization trends on the continent and increasingly globalized food chains, which Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 require better understanding of the multi-stressor context of food and livelihood security in both urban and rural contexts in Africa. [22.3.4, 22.3.4.3, 22.3.4.5] Progress has been achieved on managing risks to food production from current climate variability and near- term climate change but these will not be sufficient to address long-term impacts of climate change (high confidence). Livelihood-based approaches for managing risks to food production from multiple stressors, including rainfall variability, have increased substantially in Africa since the IPCC s Fourth Assessment Report (AR4). While these efforts can improve the resiliency of agricultural systems in Africa over the near term, current adaptations will be insufficient for managing risks from long-term climate change, which will be variable across regions and farming system types. Nonetheless, processes such as collaborative, participatory research that includes scientists and farmers, strengthening of communication systems for anticipating and responding to climate risks, and increased flexibility in livelihood options, which serve to strengthen coping strategies in agriculture for near-term risks from climate variability, provide potential pathways for strengthening adaptive capacities for climate change. [22.4.5.4, 22.4.5.7, 22.4.6, 22.6.2] Climate change may increase the burden of a range of climate-relevant health outcomes (medium confidence). Climate change is a multiplier of existing health vulnerabilities (high confidence) including insufficient access to safe water and improved sanitation, food insecurity, and limited access to health care and education. [22.3.5.1] Detection and attribution of trends is difficult because of the complexity of disease transmission, with many drivers other than weather and climate, and short and often incomplete datasets. Evidence is growing that highland areas, especially in East Africa, could experience increased malaria epidemics due to climate change (medium evidence, very high agreement). The strong seasonality of meningococcal meningitis and associations with weather and climate variability suggest the disease burden could be negatively affected by climate change (medium evidence and high agreement). The frequency of leishmaniasis epidemics in sub-Saharan Africa is changing, with spatial spread to peri-urban areas and to adjacent geographic regions, with possible contributions from changing rainfall patterns (low confidence). Climate change is projected to increase the burden of malnutrition (medium confidence), with the highest toll expected in children. [22.3.5.3] In all regions of the continent, national governments are initiating governance systems for adaptation and responding to climate change, but evolving institutional frameworks cannot yet effectively co-ordinate the range of adaptation initiatives being implemented (high confidence). Progress on national and sub-national policies and strategies has initiated the mainstreaming of adaptation into sectoral planning. [22.4.4] However, incomplete, under-resourced and fragmented institutional frameworks and overall low levels of adaptive capacity, especially competency at local government level, to manage complex socio-ecological change translate into a largely ad hoc and project-level approach, which is often donor-driven. [22.4.2, 22.4.4.3, 22.4.4.4] Overall adaptive capacity is considered to be low. [22.4.2] Disaster risk reduction, social protection, technological and infrastructural adaptation, ecosystem-based approaches and livelihood diversification are reducing vulnerability, but largely in isolated initiatives. [22.4.5] and most adaptation remain autonomous and reactive to short-term motivations. [22.4.3, 22.4.4.5] Conservation agriculture provides a viable means for strengthening resilience in agroecosystems and livelihoods that also advance adaptation goals (high confidence). A wide array of conservation agriculture practices, including agroforestry and farmer-managed natural tree regeneration, conservation tillage, contouring and terracing, and mulching are being increasingly adopted in Africa. These practices strengthen resilience of the land base to extreme events and broaden sources of livelihoods, both of which have strongly positive implications for climate risk management and adaptation. Moreover, conservation agriculture has direct adaptation-mitigation co- benefits. Addressing constraints to broader adoption of these practices, such as land tenure/usufruct stability, access to peer-to-peer learning, gender-oriented extension and credit and markets, as well as identification of perverse policy incentives would help to enable larger scale transformation of agricultural landscapes. [22.4.5.6, 22.4.5.7, 22.4.6, 22.6.2] Despite implementation limitations, Africa s adaptation experiences nonetheless highlight valuable lessons for enhancing and scaling up the adaptation response, including principles for good practice and integrated approaches to adaptation (high confidence). Five common principles for adaptation and building adaptive Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 capacity can be distilled: (i) supporting autonomous adaptation through policy that recognises the multiple stressor nature of vulnerable livelihoods; (ii) increasing attention to the cultural, ethical, and rights considerations of adaptation by increasing the participation of women, youth and poor and vulnerable people in adaptation policy and implementation; [22.4.5] (iii) combining soft path options and flexible and iterative learning approaches with technological and infrastructural approaches and blending scientific, local and indigenous knowledge when developing adaptation strategies; (iv) focusing on building resilience and implementing low-regrets adaptation with development synergies, in the face of future climate and socio-economic uncertainties; and (v) building adaptive management and social and institutional learning into adaptation processes at all levels. [22.4] Ecosystem-based approaches and pro-poor integrated adaptation-mitigation initiatives hold promise for a more sustainable and system-orientedapproach to adaptation, as does promoting equity goals, key for future resilience, through emphasising gender aspects and highly vulnerable groups such as children. . [22.4.2, 22.4.5.6, 22.6.2, Table 22-5] Strengthened inter-linkages between adaptation and development pathways and a focus on building resilience would help to counter the current adaptation deficit and reduce future maladaptation risks (high confidence). [22.4.3] Development strategies are currently not able to counter current climate risks, as highlighted by the impacts of recent extreme events; national policies that disregard cultural, traditional and context-specific factors can act as barriers to local adaptation; and there is increased knowledge of maladaptation risks from narrowly conceived development interventions and sectoral adaptation strategies that decrease resilience in other sectors or ecosystems. [22.4.4, 22.4.6] Given multiple uncertainties in the African context, successful adaptation will depend upon building resilience. [22.4, 22.5, 22.6] Options for pro-poor adaptation/resilient livelihoods include improved social protection, social services and safety nets; better water and land governance and tenure security over land and vital assets; enhanced water storage, water harvesting and post-harvest services; strengthened civil society and greater involvement in planning; and more attention to urban and peri-urban areas heavily affected by migration of poor people. [22.4.2, 22.4.4, 22.4.5, 22.4.6] Growing understanding of the multiple interlinked constraints on increasing adaptive capacity is beginning to indicate potential limits to adaptation in Africa (medium confidence). Climate change combined with other external changes (environmental, social, political, technological) may overwhelm the ability of people to cope and adapt, especially if the root causes of poverty and vulnerability are not addressed. Evidence is growing for the effectiveness of flexible and diverse development systems that are designed to reduce vulnerability, spread risk, and build adaptive capacity. These points indicate the benefits of new development trajectories that place climate resilience, ecosystem stability, equity and justice at the centre of development efforts. [22.4.6] There is increased evidence of the significant financial resources, technological support and investment in institutional and capacity development needed to address climate risk, build adaptive capacity and implement robust adaptation strategies (high confidence). Funding and technology transfer and support isto both address Africa s current adaptation deficit and to protect rural and urban livelihoods, societies and economies from climate change impacts at different local scales. [22.4,] [22.6.4] Strengthening institutional capacities and governance mechanisms to enhance the ability of national governments and scientific institutions in Africa to absorb and effectively manage large amounts of funds allocated for adaptation, will assure the effectiveness of adaptation initiatives (medium confidence). [22.6.4] Climate change and climate variability have the potential to exacerbate or multiply existing threats to human security including food, health and economic insecurity, all being of particular concern for Africa (medium confidence). [22.6.1, 22.6.1.1] Many of these threats are known drivers of conflict (high confidence). Causality between climate change and violent conflict is difficult to establish due to the presence of these and other interconnected causes, including country-specific sociopolitical, economic and cultural factors. For example, the degradation of natural resources as a result of both overexploitation and climate change will contribute to increased conflicts over the distribution of these resources. [22.6.1.1] Many of the interacting social, demographic and economic drivers of observed urbanization and migration in Africa are sensitive to climate change impacts. [22.6.1.2] A wide range of data and research gaps constrain decisionmaking in processes to reduce vulnerability, build resilience and plan and implement adaptation strategies at different levels in Africa (high confidence). Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Overarching data and research gaps identified include data management and monitoring of climate parameters and development of climate change scenarios; monitoring systems to address climate change impacts in the different sectors; research and improved methodologies to assess and quantify the impact of climate change on different sectors and systems; and socio-economic consequences of the loss of ecosystems, of economic activities, of certain mitigation choices, such as biofuels, and of adaptation strategies. [22.7] Of nine climate-related key regional risks identified for Africa, eight pose medium or higher risk even with highly adapted systems, while only one key risk assessed can be potentially reduced with high adaptation to below a medium risk level, for the end of the 21st century under 2°C global mean temperature increase above pre-industrial levels (medium confidence). Key regional risks relating to shifts in biome distribution, loss of coral reefs, reduced crop productivity, adverse effects on livestock, vector- and water-borne diseases, undernutrition, and migration are assessed as either medium or high for the present under current adaptation, reflecting Africa s existing adaptation deficit. [22.3.1, 22.3.2, 22.3.4, 22.3.5, 22.6.1.2] The assessment of significant residual impacts in a 2°C world at the end of the 21st century suggests that even under high levels of adaptation, there could be very high levels of risk for Africa. At a global mean temperature increase of 4°C, risks for Africa s food security (see key risks on livestock and crop production) are assessed as very high, with limited potential for risk reduction through adaptation. [22.3.4, 22.4.5, 22.5, Table 22-6] 22.1. Introduction Africa as a whole is one of the most vulnerable continents due to its high exposure and low adaptive capacity Climate, ecology and political boundaries in Africa vary across the continent. Since the African Union, together with its Regional Economic Communities (RECs), are encharged of the adaptation policies we have used these divisions for regional assessment within the chapter. 22.1.1. Structure of the Regions The African continent (including Madagascar) is the world s second largest and most populous continent (1,031,084,000 in 2010) behind Asia (UN DESA, 2013). The continent is organized at the regional level under the African Union (AU).1 The AU s Assembly of Heads of State and Government has officially recognized eight Regional Economic Communities (RECs) (Ruppel, 2009). Except for the Sahrawi Arab Democratic Republic,2 all AU member states are affiliated with one or more of these RECs. These RECs include the Arab Maghreb Union (AMU), with 5 countries in Northern Africa; the Community of Sahel-Saharan States (CEN-SAD), grouping 27 countries; the Common Market for Eastern and Southern Africa (COMESA), grouping 19 countries in Eastern and Southern Africa; the East African Community (EAC), with 5 countries; the Economic Community of Central African States (ECCAS), with 10 countries; the Economic Community of West African States (ECOWAS), with 15 countries; the Intergovernmental Authority on Development (IGAD) with 8 countries; and the Southern African Development Community (SADC), with 15 countries. The regional subdivision of African countries into REC s is a structure used by the AU and the New Partnership for Africa (NEPAD). [FOOTNOTE 1: Due to the controversies regarding the Sahrawi Arab Democratic Republic, Morocco withdrew from the Organization of African Unity (OAU) in protest in 1984 and, since South Africa's admittance in 1994, remains the only African nation not within what is now the AU.] [FOOTNOTE 2: Although the Sahrawi Arab Democratic Republic has been a full member of the OAU since 1984 and remains a member of the AU, the republic is not generally recognized as a sovereign state and has no representation in the UN.] Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.1.2. Major Conclusions from Previous Assessments 22.1.2.1. Regional Special Report and Assessment Reports Refer to Table 22-1 for a brief summary of conclusions from previous IPCC assessments. [INSERT TABLE 22-1 HERE Table 22-1: Major conclusions from previous IPCC assessments.] 22.1.2.2. Special Report on Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX) The Special Report of the IPCC on managing the risks of extreme events and disasters to advance climate change adaptation (IPCC, 2012) is of particular relevance to the African continent. There is low to medium confidence in historical extreme temperature and heavy rainfall trends over most of Africa because of partial lack of data, literature and lack of consistency of reported patterns in the literature (Seneviratne et al., 2012). However, most regions within Africa for which data is available have recorded an increase in extreme temperatures (Seneviratne et al., 2012). For projected temperature extreme there is high confidence that heat waves and warm spell durations will increase, suggesting an increased persistence of hot days (90th percentile) toward the end of the century (Tebaldi et al., 2006; Orlowsky and Seneviratne, 2012). There is high confidence for projected shorter extreme maximum temperature return periods across the B1, A1B and A2 scenarios for the near and far future as well as a reduction of the number of cold extremes (Seneviratne et al., 2012). In East and southern Africa, there is medium confidence that droughts will intensify in the 21st Century in some seasons, due to reduced precipitation and/or increased evapotranspiration. There is low confidence in projected increases of heavy precipitation over most of Africa except over East Africa where there is a high confidence in a projected increase in heavy precipitation (Seneviratne et al., 2012). 22.2. Observed Climate Trends and Future Projections 22.2.1. Temperature 22.2.1.1. Observed Trends Near surface temperatures have increased by 0.5°C or more during the last 50-100 years over most parts of Africa with minimum temperatures warming more rapidly than maximum temperatures (Hulme et al., 2001; Jones and Moberg, 2003; Kruger and Shongwe, 2004; Schreck and Semazzi, 2004; New et al., 2006; IPCC, 2007; Rosenzweig et al., 2007; Trentberth et al., 2007; Christy et al., 2009; Collins 2011; Grab and Craparo, 2011; Hoffman et al., 2011; Mohamed, 2011; Stern et al. 2011; Funk et al., 2012; Nicholson et al., 2013). Near surface air temperature anomalies in Africa were significantly higher for the period 1995 2010 compared to the period 1979 1994 (Collins, 2011). Figure 22-1 shows that it is very likely that mean annual temperature has increased over the past century over most of the African continent, with the exception of areas of the interior of the continent where the data coverage has been determined to be insufficient to draw conclusions about temperature trends (Figure 22-1, Box CC-RC). There is strong evidence of an anthropogenic signal in continent-wide temperature increases in the 20th century (WGI 10.3.1; Stott, 2003; Min and Hense, 2007, Stott et al., 2010; Stott et al., 2011). In recent decades North African annual and seasonal observed trends in mean near surface temperature indicates an overall warming that is significantly beyond the range of changes due to natural (internal) variability (Barkhordarian et al., 2012a). During the warm seasons (March-April-May, June-July-August) an increase in near surface temperature is shown over north Algeria and Morocco which is very unlikely due to natural variability or natural forcing alone (Barkhordarian et al., 2012b). The region has also experienced positive trends in annual minimum and maximum temperature (Vizy and Cook, 2012). Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Over West Africa and the Sahel near surface temperatures have increased over the last 50 years. Using indices developed by the Expert Team on Climate Change Detection and Indices (ETCCDI), New et al. (2006) show the number of cold days and cold nights have decreased and the number of warm days and warm nights have increased between 1961 and 2000. Many of these trends are statistically significant at the 90% level and they find similar trends in extreme temperature indices. Collins (2011) shows statistically significant warming of between 0.5-0.8 degrees between 1970 and 2010 over the region using remotely sensed data with a greater magnitude of change in the latter 20 years of the period compared to the former. The equatorial and southern parts of eastern Africa have experienced a significant increase in temperature since the beginning of the early 1980s (Anyah and Qiu, 2012). Similarly, recent reports from the Famine Early Warning Systems Network (FEWS NET) indicate that there has been an increase in seasonal mean temperature in many areas of Ethiopia, Kenya, South Sudan, and Uganda over the last 50 years (Funk et al., 2011, 2012). In addition, warming of the near surface temperature and an increase in the frequency of extreme warm events has been observed for countries bordering the western Indian Ocean between 1961 and 2008 (Vincent et al., 2011b). In recent decades, most of southern Africa has also experienced upward trends in annual mean, maximum, and minimum temperature over large extents of the subregion during the last half of the 20th century, with the most significant warming occurring during the last two decades (Zhou et al., 2010; Collins, 2011; Kruger and Sekele, 2012). Minimum temperatures have increased more rapidly relative to maximum temperatures over inland southern Africa (New et al., 2006). 22.2.1.2. Projected Trends Temperatures in Africa are projected to rise faster than the global average increase during the 21st Century (Christensen et al., 2007; Joshi et al. 2011; Sanderson et al., 2011; James and Washington, 2013). Global average near-surface air temperature are projected to move beyond 20th Century simulated variability by 2069 (+/-18 years) under RCP4.5 and by 2047 (+/-14 years) under RCP8.5 (Mora et al. 2013). However, in the tropics, especially tropical West Africa, these unprecedented climates are projected to occur one to two decades earlier the global average because the relatively small natural climate variability in this region generates narrow climate bounds that can be easily surpassed by relatively small climate changes. Figure 22-1 shows projected temperature increases based on the CMIP5 ensemble. Increases in mean annual temperature over all land areas are very likely in the mid- and late-21st-century periods for RCP2.6 and RCP8.5 (Figure 22-1, Box CC-RC). Ensemble-mean changes in mean annual temperature exceed 2C above the late-20th-century baseline over most land areas of the continent in the mid- 21st-century for RCP8.5, and exceed 4C over most land areas in the late-21st-century for RCP8.5. Changes in mean annual temperature for RCP8.5 follow a pattern of larger changes in magnitude over northern and southern Africa, with (relatively) smaller changes in magnitude over central Africa. The ensemble-mean changes are less than 2C above the late-20th-century baseline in both the mid- and late-21st-century for RCP2.6. Over North Africa under the A1B scenario, both annual minimum and maximum temperature are likely to increase in the future, with greater increase in minimum temperature (Vizy and Cook, 2012). The faster increase in minimum temperature is consistent with greater warming at night, resulting in a decrease in the future extreme temperature range (Vizy and Cook, 2012). Higher temperature increases are projected during boreal summer by CMIP5 GCMs (WGI Annex 1). A strengthening of the North African thermal low in 21st century is associated with a surface temperature increase (Paeth et al., 2009; Patricola and Cook, 2010; Barkhordarian et al., 2012a; Cook and Vizy, 2012). Temperature projections over West Africa for the end of the 21st Century from both the CMIP3 GCMs (A2 and A1B scenarios) and CMIP5 GCMs (RCP4.5 and RCP8.5) range between 3-6°C above the late 20th Century baseline (Meehl et al. 2007; Fontaine et al. 2011; Diallo et al., 2012; Monrie et al. 2012; Figure 22-1; Figure 22-2). Regional downscalings over the region produce a similar range of projected change (Patricola and Cook, 2010; Mariotti et al., 2011; Patricola and Cook, 2011; Vizy et al., 2012). Diffenbaugh and Giorgi (2012) identify the Sahel and tropical West Africa as a hotspot of climate change for both RCP4.5 and RCP8.5 pathways and unprecedented climates are projected to occur earliest (late 2030s to early 2040s) in these regions (Mora et al., 2013). Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climate model projections under the A2 and B1 scenarios over Ethiopia show warming in all four seasons across the country, which may cause a higher frequency of heat waves as well as higher rates of evaporation (Conway and Schipper, 2011). Projected maximum and minimum temperatures over equatorial eastern Africa show a significant increase in the number of days warmer than 2°C above the 1981 2000 average by the middle and end of the 21st century under the A1B and A2 scenarios (Anyah and Qiu, 2012). Elshamy et al. (2009) show a temperature increase over the upper Blue Nile of between 2°C and 5°C at the end of the 21st Century under the A1B scenario compared to a 1961-1990 baseline. Mean land surface warming in Southern Africa is likely to exceed the global mean land surface temperature increase in all seasons (Sillmann and Roeckner, 2008; Watterson, 2009; Mariotti et al., 2011; James and Washington, 2013; Orlowsky and Seneviratne, 2012). Furthermore, towards the end of the 21st Century the projected warming of between 3.4-4.2C above the 1981-2000 average under the A2 scenario far exceeds natural climate variability (Moise and Hudson, 2008). High warming rates are projected over the semi-arid southwestern parts of the subregion covering northwestern South Africa, Botswana, and Namibia (WGI Annex 1; Moise and Hudson, 2008; Engelbrecht et al., 2009; Watterson, 2009; Shongwe et al., 2009). Observed and simulated variations in past and projected future annual average temperature over five African regions (UMA, SADC, ECCAS, ECOWAS, and COMESA) are captured in Figure 22-2 which indicates the projected temperature rise is very likely to exceed the 1986-2005 baseline by between 3 and 6 °C across these regions by the end of this century under RCP8.5. [INSERT FIGURE 22-1 HERE Figure 22-1: Observed and simulated variations in past and projected future annual average precipitation and temperature. Observed differences in the Climate Research Unit, University of East Anglia data (CRU) are shown between the 1986-2005 and 1906-1925 periods, with white indicating areas where the difference between the 1986- 2005 and 1906-1925 periods is less than twice the standard deviation of the 20 20-year periods beginning in the years 1906 through 1925. For CMIP5, white indicates areas where <66% of models exhibit a change greater than twice the baseline standard deviation of the respective model s 20 20-year periods ending in years 1986 through 2005. Grey indicates areas where >66% of models exhibit a change greater than twice the respective model baseline standard deviation, but <66% of models agree on the sign of change. Colors with circles indicate the ensemble-mean change in areas where >66% of models exhibit a change greater than twice the respective model baseline standard deviation and >66% of models agree on the sign of change. Colors without circles indicate areas where >90% of models exhibit a change greater than twice the respective model baseline standard deviation and >90% of models agree on the sign of change. The realizations from each model are first averaged to create baseline-period and future-period mean and standard deviation for each model, from which the multi-model mean and the individual model signal-to-noise ratios are calculated. The baseline period is 1986-2005. The late-21st Century period is 2081- 2100. The mid-21st century period is 2046-2065.] [INSERT FIGURE 22-2 HERE Figure 22-2: Observed and simulated variations in past and projected future annual average temperature over EAC- IGAD-Egypt, ECCAS, ECOWAS, SADC and UMA. Black lines show various estimates from observational measurements. Shading denotes the 5-95 percentile range of climate model simulations driven with "historical" changes in anthropogenic and natural drivers (63 simulations), historical changes in "natural" drivers only (34), the "RCP2.6" emissions scenario (63), and the "RCP8.5" (63). Data are anomalies from the 1986-2005 average of the individual observational data (for the observational time series) or of the corresponding historical all-forcing simulations. Further details are given in Box 21-3.] 22.2.2. Precipitation 22.2.2.1. Observed Changes Most areas of the African continent lack sufficient observational data to draw conclusions about trends in annual precipitation over the past century (Figure 22-1, Box CC-RC). Additionally, in many regions of the continent discrepancies exist between different observed precipitation datasets (Nikulin et al. 2012; Sylla et al., 2012; Kim et Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 al. 2013; Kalognomou et al., 2013). Areas where there are sufficient data include very likely decreases in annual precipitation over the past century over parts of the western and eastern Sahel region in northern Africa, along with very likely increases over parts of eastern and southern Africa. Over the last few decades the northern regions of North Africa (north of the Atlas Mountains and along the Mediterranean coast of Algeria and Tunisia) have experienced a strong decrease in the amount of precipitation received in winter and early spring (Barkhordarian et al., 2013). The observed record also indicates greater than 330 dry days (with less than 1 mm/day rainfall) per year over the 1997 - 2008 time period (Vizy and Cook, 2012). However, in autumn (September, October, November) observations show a positive trend in precipitation in some parts of North Algeria and Morocco (Barkhordarian et al., 2013). The Sahara Desert, which receives less than 25 mm/year, shows little seasonal change (Liebmann et al., 2012). Rainfall over the Sahel has experienced an overall reduction over the course of the 20th Century with a recovery toward the last 20 years of the century (WGI 14.3.7.1; Nicholson et al., 2000; Lebel and Ali, 2009; Ackerley et al., 2011; Mohamed, 2011; Biasutti, 2013). The occurrence of a large number of droughts in the Sahel during the 1970s and 1980s is well documented and understood (Biasutti and Giannini, 2006; Biasutti et al., 2008; Greene et al., 2009). The recovery of the rains may be due to natural variability (Mohino et al., 2011) or a forced response to increased greenhouse gases (Haarsma et al., 2005; Biasutti, 2013) or reduced aerosols (Ackerley et al., 2011). Precipitation in eastern Africa shows a high degree of temporal and spatial variability dominated by a variety of physical processes (Rosell and Holmer, 2007; Hession and Moore, 2011). Williams and Funk (2011) and Funk et al. (2008) indicate that over the last three decades rainfall has decreased over eastern Africa between March and May/June. The suggested physical link to the decrease in rainfall is the rapid warming of Indian Ocean, which causes an increase in convection and precipitation over the tropical Indian Ocean and thus contributes to increased subsidence over eastern Africa and a decrease in rainfall during March to May/June (Funk et al., 2008; Williams and Funk, 2011). Similarly, Lyon and DeWitt (2012) show a decline in the March May seasonal rainfall over eastern Africa. Summer (June September) monsoonal precipitation has declined throughout much of the Great Horn of Africa over the last 60 years [during the 1948 2009 period; Williams et al., (2012)] as a result of the changing sea level pressure (SLP) gradient between Sudan and the southern coast of the Mediterranean Sea and the southern tropical Indian Ocean region (Williams et al., 2012). Over Southern Africa a reduction in late austral summer precipitation has been reported over its western parts, extending from Namibia, through Angola, and toward the Congo during the second half of the 20th Century (Hoerling et al., 2006; New et al., 2006). The drying is associated with an upward trend in tropical Indian Ocean Sea Surface Temperatures (SSTs). Modest downward trends in rainfall are found in Botswana, Zimbabwe, and western South Africa. Apart from changes in total or mean summer rainfall, certain intra-seasonal characteristics of seasonal rainfall such as onset, duration, dry spell frequencies, and rainfall intensity as well as delay of rainfall onset have changed (Tadross et al., 2005; Thomas et al., 2007; Tadross et al., 2009; Kniveton et al., 2009). An increasing frequency of dry spells is accompanied by an increasing trend in daily rainfall intensity which has implications for run-off characteristics (New et al., 2006). 22.2.2.2. Projected Changes Precipitation projections are more uncertain than temperature projections (Rowell, 2012) and exhibit higher spatial and seasonal dependence than temperature projections (Orlowsky and Seneviratne, 2012). The CMIP5 ensemble projects very likely decreases in mean annual precipitation over the Mediterranean region of northern Africa in the mid- and late-21st century periods for RCP8.5 (Figure 22-1, Box CC-RC). CMIP5 also projects very likely decreases in mean annual precipitation over areas of southern Africa beginning in the mid-21st-century for RCP8.5 and expanding substantially in the late-21st-century for RCP8.5. In contrast, CMIP5 projects likely increases in mean annual precipitation over areas of central and eastern Africa beginning the mid-21st-century for RCP8.5. Most areas of the African continent do not exhibit changes in mean annual precipitation that exceed the baseline variability in more than 66% of the models in either the mid- or late-21st-century periods for RCP2.6. Observed and simulated Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 variations in past and projected future annual average precipitation over five African regions (UMA, ECCAS, ECOWAS, SADC and COMESA) are captured in Figure 22-2. A reduction in rainfall over northern Africa is very likely by the end of the 21st Century. The annual and seasonal drying/warming signal over the Northern African region (including North of Morocco, Algeria, Libya, Egypt and Tunisia) is a consistent feature in the global (Giorgi and Lionello, 2008; Barkhordarian et al., 2013) and the regional (Lionello and Giorgi, 2007; Gao and Giorgi, 2008; Paeth et al., 2009; Patricola and Cook, 2010) climate change projections for the 21st Century under the A1B and A2 scenarios. Furthermore, over the northern basin of Tunisia, climate models under the A1B scenario project a significant decrease in the median, and 10th and 90th percentiles values of precipitation in winter and spring seasons (Bargaoui et al., 2013). West African precipitation projections in the CMIP-3 and CMIP-5 archives show inter-model variation in both the amplitude and direction of change that is partially attributed to the inability of GCMs to resolve convective rainfall (WGI 14.8.7; Biasutti et al., 2008; Druyan, 2011; Fontaine et al., 2011; Roehrig et al., 2013). Many CMIP-5 models indicate a wetter core rainfall season with a small delay to rainy season by the end of the 21st Century ( WGI 14.8.7; Biasutti, 2013). However, regional climate models (RCMs) can alter the sign of rainfall change of the driving GCM especially in regions of high or complex topography (WGI 14.3.7.1; WGI 9.6.4; Sylla et al., 2012; Cook and Vizy, 2013; Saeed et al. 2013). There is therefore low to medium confidence in the robustness of projected regional precipitation change until a larger body of regional results become available through, for example, the coordinated regional downscaling experiment, CORDEX (Giorgi et al., 2009, Jones et al., 2011, Hewitson et al., 2012). An assessment of 12 CMIP-3 GCMs over eastern Africa suggest that by the end of the 21st Century there will be a wetter climate with more intense wet seasons and less severe droughts during October-November-December (OND) and March-April-May (MAM) (WGI 14.8.7; Moise and Hudson, 2008; Shongwe et al., 2011). These results indicate a reversal of historical trend in these months (Williams and Funk, 2011; Funk et al. 2008). Lyon and DeWitt (2012) ascribe this reversal to recent cooling in the eastern equatorial Pacific that offsets the equatorial Pacific SST warming projected by CMIP3 GCMs in future scenarios. However, GCM projections over Ethiopia indicate a wide range of rainfall spatial pattern changes (Conway and Schipper, 2011) and in some regions GCMs do not agree on the direction of precipitation change, e.g. in the upper Blue Nile basin in the late 21st Century (Elshamy et al., 2009). Regional climate model studies suggest drying over most parts of Uganda, Kenya, and South Sudan in August and September by the end of the 21st Century as a result of a weakening Somali jet and Indian monsoon (Patricola and Cook, 2011). Cook and Vizy (2013) indicate truncated boreal spring rains in the mid-21st Century over eastern Ethiopia, Somalia, Tanzania and southern Kenya while the boreal fall season is lengthened in the southern Kenya and Tanzania (Nakaegawa et al., 2012). These regional studies highlight the importance of resolving both regional scale atmospheric processes and local effects like land surface on rainfall simulation across the region (WGI 14.8.7). Over southern Africa CMIP3 GCM projections show a drying signal in the annual mean over the climatologically dry southwest, extending northeastward from the desert areas in Namibia and Botswana (Moise and Hudson, 2008; James and Washington, 2013; Orlowsky and Seneviratne, 2012). This pattern is replicated by CMIP5 GCMs (see Figure 22-1). During the austral summer months, dry conditions are projected in the southwest while downscaled projections indicate wetter conditions in the southeast of South Africa and the Drakensberg mountain range (Hewitson and Crane, 2006; Engelbrecht et al., 2009). Consistent with the AR4, drier winters are projected over a large area in southern Africa by the end of the century as a result of the poleward displacement of mid-latitude storm tracks (WGI 14.8.7; Moise and Hudson, 2008; Engelbrecht et al., 2009; Shongwe et al., 2009; Seth et al., 2011; James and Washington, 2013). Rainfall decreases are also projected during austral spring months, implying a delay in the onset of seasonal rains over a large part of the summer rainfall region of Southern Africa (Shongwe et al., 2009; Seth et al., 2011). The sign, magnitude and spatial extent of projected precipitation changes are dependent on the coupled general circulation model (CGCM) employed, due primarily to parameterization schemes used and their interaction with model dynamics (Hewitson and Crane, 2006; Rocha et al., 2008). Changes in the parameterization schemes of a single regional climate model produced opposite rainfall biases over the region (Crétat et al., 2012) so multiple ensemble downscalings, such as those being produced through CORDEX, are important to more fully describe the uncertainty associated with projected rainfall changes across the African continent (WGI 9.6.5; Laprise et al., 2013). Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.2.3. Observed and Projected Changes in Extreme Temperature and Rainfall In Northern Africa the north western Sahara experienced 40-50 heat wave days per year during the 1989-2009 time period (Vizy and Cook, 2012). There is a projected increase in this number of heat wave days over the 21st century (Patricola and Cook, 2010; Vizy and Cook, 2012). Over West Africa there is low to medium confidence in projected changes of heavy precipitation by the end of the 21st Century based on CMIP-3 GCMs (Seneviratne et al., 2012). Regional model studies suggest an increase in the number of extreme rainfall days over West Africa and the Sahel during May and July (Vizy and Cook, 2012) and more intense and more frequent occurrences of extreme rainfall over the Guinea Highlands and Cameroun Mountains (Sylla et al., 2012; Haensler et al., 2013. The ability of regional climate models to resolve complex topography captures the amplifying role of topography in producing extreme rainfall that GCMs cannot. Extreme precipitation changes over Eastern Africa such as droughts and heavy rainfall have been experienced more frequently during the last 30-60 years (Funk et al., 2008; Williams and Funk, 2011; Shongwe et al., 2011; Lyon and DeWitt, 2012). A continued warming in the Indian Pacific warm pool has been shown to contribute to more frequent East African droughts over the past 30 years during the spring and summer seasons (Williams and Funk, 2011). It is unclear whether these changes are due to anthropogenic influences or multidecadal natural variability (Lyon and DeWitt, 2012; Lyon et al., 2013). Projected increases in heavy precipitation over the region have been reported with high certainty in the SREX (Seneviratne et al., 2012) and Vizy and Cook (2012) indicate an increase in the number of extreme wet days by the mid-20th Century. Over southern Africa an increase in extreme warm ETCCDI indices (hot days, hot nights, hottest days) and a decrease in extreme cold indices (cold days and cold nights) in recent decades is consistent with the general warming trend (New et al., 2006; Tebaldi et al., 2006; Aguilar et al. 2009; Kruger and Sekele, 2012). The probability of austral summer heat waves over South Africa increased over the last two decades of the 20th century compared to 1961 to 1980 (Lyon, 2009). Enhanced heat wave probabilities are associated with deficient rainfall conditions that tend to occur during El Nino events. The southwestern regions are projected to be at a high risk to severe droughts during the 21st Century and beyond (Hoerling et al., 2006; Shongwe et al., 2011). Large uncertainties surround projected changes in tropical cyclone landfall from the southwest Indian Ocean that have resulted in intense floods during the 20th Century. Future precipitation projections show changes in the scale of the rainfall probability distribution, indicating that extremes of both signs may become more frequent in the future (Kay and Washington, 2008). 22.3. Vulnerability and Impacts This section highlights Africa s vulnerability to climate change, as well as the main observed and potential impacts on natural resources, ecosystems, and economic sectors. Figure 22-3 summarizes the main conclusions regarding observed changes in regional climate and their relation to anthropogenic climate change (described in 22.2) as well as regarding observed changes in natural and human systems and their relation to observed regional climate change (described in this section). Confidence in detection and attribution of anthropogenically-driven climate change is highest for temperature measures. In many regions of Africa, evidence is constrained by limited monitoring. However, impacts of observed precipitation changes are amongst the observed impacts with the highest assessment of confidence, implying that some of the potentially more significant impacts of anthropogenic climate change for Africa are of a nature that challenges detection and attribution analysis (18.5.1). [INSERT FIGURE 22-3 HERE Left: Confidence in detection and in attribution of observed climate change over Africa to anthropogenic emissions. All detection assessments are against a reference of no change, while all attribution assessments concern a major role of anthropogenic emissions in the observed changes. See 22.2, and SREX-3, and WGI AR5 10 for details. Right: Confidence in detection and in attribution of the impacts of observed regional climate change on various African systems. All detection assessments are against a reference of no change, except "Kenyan Highlands malaria" Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (changes due to vaccination, drug resistance, demography, and livelihoods), "Great Lakes fisheries" (changes due to fisheries management and land use) and Adapting South African farmers (economic changes). Attribution is to a major role or a minor role of observed climate change, as indicated. See 22.2.2, 22.3.2.1, 22.3.2.2, 2.3.3, 22.4.2, 22.3.4.4, 22.3.5.4, 22.4.5.7 and Tables 18-5, 18-6, 18-7, and 18-9 for details. Assessments follow the methods outlined in 18.2.] 22.3.1. Socioeconomic and Environmental Context Influencing Vulnerability and Adaptive Capacity Equitable socioeconomic development in Africa may strengthen its resilience to various external shocks, including climate change. In 2009, the Human Rights Council adopted Resolution 10/43 which noted the effects of climate change on the enjoyment of human rights, and reaffirmed the potential of human rights obligations and commitments to inform and strengthen international and national policy making. [FOOTNOTE 3: U.N. Doc. A/HRC/10/L.11.] The impacts of climate change on human rights have been explicitly recognised by the African Commission on Human and Peoples Rights (hereafter African Commission) in its Resolution on Climate Change and Human Rights and the Need to Study its Impact in Africa (ACHPR/Res 153 XLV09). The 1981 African (Banjul) Charter on Human and Peoples Rights (hereafter African Charter) protects the right of peoples to a general satisfactory environment favorable to their development (Article 24). The recognition of this right and the progressive jurisprudence by the African Commission in environmental matters underline the relevance of potential linkages between climate change and human rights (Ruppel, 2012). The link between climate change and humans is not only associated with human rights. Rather, strong links exist between climate change and the MDGs: climate change may adversely affect progress toward attaining the MDGs, as climate change can increase the pressure not only economic activities, such as agriculture (22.3.4) and fishing (22.3.4.4) but also adversely affect urban areas located in coastal zones (22.3.6). Slow progress in attaining most MDGs may, meanwhile, reduce the resilience and adaptive capabilities of African individuals, communities, states, and nations (ECA et al., 2009, 2012; UNDP et al., 2011). The African continent has made significant progress on some MDGs; however, not all MDGs have been achieved, yet with high levels of spatial and group disparities. Additionally, progress on all MDG indicators is skewed in favor of higher-income groups and urban populations, which means further marginalization of already excluded groups (UN et al., 2008; AfDB et al., 2010; World Bank and IMF, 2010). As a whole, the continent is experiencing a number of demographic and economic constraints, with the population having more than doubled since 1980; exceeding one billion in 2010 and expected to reach three billion by the year 2050, should fertility remain constant (Muchena et al., 2005; Fermont et al., 2008; UN DESA, 2011). The global economic crisis is adding additional constraints on economic development efforts leading to increased loss of livelihood and widespread poverty (Moyo, 2009; Easterly, 2009; Adesina, 2010). The percent of the population below the poverty line has decreased from 56.5% in 1990 to 47.5% in 2008 (excluding North Africa); however a significant proportion of the population living below the poverty line remains chronically poor (ECA et al., 2012). Although poverty in rural areas in Sub-Saharan Africa has declined from 64.9% in 1998 to 61.6% in 2008, it is still double the prevailing average in developing countries in other regions (IFAD, 2010). Agriculture, which is the main economic activity in terms of employment share, is 98% rain fed in the sub-Saharan region (FAO, 2002). 4 Stagnant agricultural yields, relative to the region s population growth, have led to a fall in per capita food availability since the 1970s (UN et al., 2008).5 Such stagnation was reversed with an improved performance of the agricultural sector in sub Saharan Africa during the 2000 2010. However, most of this improvement was the result of countries recovering from the poor performance of the 1980s and 1990s along with favorable domestic prices (Nin-Pratt et al., 2012). In addition, recent increases in global food prices aggravate food insecurity among the urban poor, increasing the risk of malnutrition and its consequences (UN et al., 2008). For example, it was estimated that the global rise in food Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 prices has contributed to the deaths of an additional 30,000 to 50,000 children suffering from malnutrition in 2009 in sub-Saharan Africa (Friedman and Schady, 2009) see Table 22.2. This situation may be complicated further by changes in rainfall variability and extreme weather events affecting the agriculture sector (Yabi and Afouda, 2012). In response, the New Partnership for Africa s Development (NEPAD) was founded in 2001, for Africans to take the lead in efforts to achieve the development vision espoused in the AU Constitutive Act as well as the MDGs and to support regional integration as a mechanism for inclusive growth and development in Africa (NEPAD et al., 2012). Furthermore, the Comprehensive Africa Agriculture Development Program (CAADP), which works under the umbrella of NEPAD, was established in 2003 to help African countries reach a higher path of economic growth through agriculture-led development. For this to happen, it focuses on four pillars for action: land and water management, market access, food supply and hunger, and agricultural research (NEPAD, 2010). [INSERT TABLE 22-2 HERE Table 22-2: Under-nourishment in Africa, by number and % of total population.] [FOOTNOTE 4: However, mining and energy sectors, where active, are undergoing expansion, stimulating growth and adding potentially to state revenues but are also highly vulnerable to global recession. overall, the limited production and export structures of the continent are likely to maintain its historical vulnerability to external shocks (ECA and AUC, 2011).] [FOOTNOTE 5: Lack of extension services for farmers in Africa can also contribute to low utilization and spread of innovations and technologies that can help mitigate climate change.] Africa has made much progress in the achievement of universal primary education; however, the results are unevenly distributed. Nevertheless, a considerable number of children, especially girls from poor backgrounds and rural communities, still do not have access to primary education (UN et al., 2008). From the livelihood perspective, African women are vulnerable to the impacts of climate change because they shoulder an enormous but imprecisely recorded portion of the responsibility for subsistence agriculture, the productivity of which can be expected to be adversely affected by climate change and over-exploited soil (Viatte et al., 2009; see also 22.4.2 and Table 22-5).6 Global financial crises, such as the one experienced in 2007/2008, as well as downturn economic trends at national level, may cause job losses in the formal sector and men may compete for jobs in the informal sector that were previously undertaken by women, making them more vulnerable (AfDB et al., 2010). [FOOTNOTE 6: For instance, 84% of women in sub-Saharan Africa, compared with 69.5% of men, are engaged in such jobs. In Northern Africa, even though informal or self-employment is less predominant, the gender gap is stark, with much higher proportion of women compared to men are in the more vulnerable informal and self-employed status (56.7% of women compared with 34.9% of men) (UN DESA, 2011).] Significant efforts have been made to improve access to safe drinking water and sanitation in Africa, with access to safe drinking water increasing from 56 to 65% between 1990 and 2008 (UNDP et al., 2011), with sub-Saharan Africa nearly doubling the number of people using an improved drinking water source from 252 to 492 million over the same period (UN, 2011). Despite such progress, significant disparities in access to safe water and sanitation, between not only urban and rural but also between large- and medium- and small-sized cities, still exist (UNDP et al., 2011). Use of improved sanitation facilities, meanwhile, is generally low in Africa, reaching 41% in 2010 compared to 36% in 1990 (UNDP et al., 2011). 22.3.2. Ecosystems It is recognized that interactions between the different drivers of ecosystem structure, composition, and function are complex, which makes the prediction of the impacts of climate change more difficult (see Chapter 4). In AR4, the chapter on Africa indicated that extensive pressure is exerted on different ecosystems by human activities Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (deforestation, forest degradation, biomass utilisation for energy) as well as processes inducing changes such as fires or desertification (see WGII AR4 9.2.2.7). Even if the trend is toward better preservation of ecosystems and a decrease in degradation (like deforestation), pressures linked, for example, to agriculture and food security, energy demand, and urbanization are increasing, putting these ecosystems at risk. This chapter emphasizes new information since AR4 regarding the vulnerability to and impacts of climate change for some terrestrial, fresh water and coastal/ocean ecosystems. 22.3.2.1. Terrestrial Ecosystems Changes are occurring in the distribution and dynamics of all types of terrestrial ecosystems in Africa, including deserts, grasslands and shrublands, savannas and woodlands, and forests (high confidence) (see also 4.3.2.5). Since AR4, three primary trends have been observed at the continental scale. The first is a small overall expansion of desert and contraction of the total vegetated area (Brink and Eva, 2009) (low confidence). The second is a large increase in the extent of human influence within the vegetated area, accompanied by a decrease in the extent of natural vegetation (Brink and Eva, 2009; Potapov et al., 2012; Mayaux et al., 2013) (high confidence). The third is a complex set of shifts in the spatial distribution of the remaining natural vegetation types, with net decreases in woody vegetation in western Africa (Vincke et al., 2010; Ruelland et al., 2011; Gonzalez et al., 2012) and net increases in woody vegetation in central, eastern, and southern Africa (Wigley et al., 2009; Wigley et al., 2010; Buitenwerf et al., 2012; Mitchard and Flintrop, 2013) (high confidence). Overall, the primary driver of these changes is anthropogenic land use change, particularly the expansion of agriculture, livestock grazing, and fuelwood harvesting (Brink and Eva, 2009; Kutsch et al., 2011; Bond and Midgley, 2012; Gonzalez et al., 2012) (high confidence). Natural climate variability, anthropogenic climate change, and interactions between these drivers and anthropogenic land use change have important additional and interacting effects (Foden et al., 2007; Touchan et al., 2008; Brink and Eva, 2009; Bond and Midgley, 2012; Gonzalez et al., 2012) (high confidence). Due to these interactions, it has been difficult to determine the role of climate change in isolation from the other drivers (Malhi et al., 2013). In general, while there are already many examples of changes in terrestrial ecosystems that are consistent with a climate change signal and have been detected with high confidence, attribution to climate change has tended to be characterized by low confidence [see Table 22-3]. New observations and approaches are improving confidence in attribution (e.g., Buitenwerf et al., 2012, Gonzalez et al., 2012, Pettorelli et al., 2012, Otto et al., 2013). There is high agreement that continuing changes in precipitation, temperature, and CO2 associated with climate change are very likely to drive important future changes in terrestrial ecosystems throughout Africa (high confidence) [see examples in 4.3.3.1, 4.3.3.2]. Modeling studies focusing on vegetation responses to climate have projected a variety of biome shifts, primarily related to the extent of woody vegetation (Delire et al., 2008; Gonzalez et al., 2010; Bergengren et al., 2011; Zelazowski et al., 2011; Midgley, 2013). For an example of such projections, see Figure 22-4. However, substantial uncertainties are inherent in these projections because vegetation across much of the continent is not deterministically driven by climate alone (high confidence). Advances in understanding how vegetation dynamics are affected by fire, grazing, and the interaction of fire and grazing with climate are expected to enable more sophisticated representations of these processes in coupled models (Scheiter and Higgins, 2009; Staver et al., 2011a; Staver et al., 2011b). Improvements in forecasting vegetation responses to climate change should reduce the uncertainties that are currently associated with vegetation feedbacks to climate forcing, as well as the uncertainties about impacts on water resources, agriculture, and health (Alo and Wang, 2008; Sitch et al., 2008) [see 4.5]. [INSERT FIGURE 22-4 HERE Figure 22-4: Left Projected biome change from the periods 1961-1990 to 2071-2100 using the MC1 Dynamic Vegetation Model. Change is indicated if any of nine combinations of three GCMs (CSIRO Mk3, HadCM3, MIROC 3.2 medres) and three emissions scenarios (B1, A1B, A2) project change and is thus a worst-case scenario. Colours represent the future biome predicted. Right Vulnerability of ecosystems to biome shifts based on historical climate (1901-2002) and projected vegetation (2071-2100), where all nine GCM-emissions scenario combinations agree on the projected biome change. Source: Gonzalez et al. (2010).] Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.3.2.2. Freshwater Ecosystems Freshwater ecosystems in Africa are at risk from anthropogenic land use change, over-extraction of water and diversions from rivers and lakes, and increased pollution and sedimentation loading in water bodies (Vörösmarty et al., 2005; Vié et al., 2009; Darwall et al., 2011). Climate change is also beginning to affect freshwater ecosystems (see also Chapter 3.5.2.4, as evident by elevated water temperatures reported in surface waters of Lakes Kariba, Kivu, Tanganyika, Victoria, and Malawi (Odada et al., 2006; Verburg and Hecky, 2009; Marshall et al., 2009; Hecky et al., 2010; Magadza, 2010; Tierney et al., 2010; Olaka et al., 2010; Magadza, 2011; Ndebele-Murisa, 2011; Woltering et al., 2011; Osborne, 2012; Ndebele-Murisa et al., 2012) (medium confidence). Small variations in climate cause wide fluctuations in the thermal dynamics of freshwaters (Odada et al., 2006; Stenuite et al., 2007; Verburg and Hecky, 2009; Moss, 2010; Olaka et al., 2010). Thermal stratification in the regions lakes, for instance, isolates nutrients from the euphotic zone, and is strongly linked to hydrodynamic and climatic conditions (Sarmento et al., 2006; Ndebele-Murisa et al., 2010). Moderate warming may be contributing to reduced lake water inflows and therefore nutrients, which subsequently destabilizes plankton dynamics and thereby adversely affects food resources for higher trophic levels of mainly planktivorous fish (low confidence) (Magadza, 2008; Verburg and Hecky, 2009; Magadza, 2010; Ndebele-Murisa et al., 2011). However, the interacting drivers of fisheries decline in African lakes are uncertain, given the extent to which other factors, such as overfishing, pollution, and invasive species also impact lake ecosystems and fisheries production (Phoon et al., 2004; Sarvala et al., 2006; Verburg et al., 2007; Tumbare, 2008; Hecky et al., 2010; Marshall, 2012). 22.3.2.3. Coastal and Ocean Systems Coastal and ocean systems are important for the economies and livelihoods of African countries, and climate change will increase challenges from existing stressors, such as overexploitation of resources, habitat degradation, loss of biodiversity, salinization, pollution, and coastal erosion (Arthurton et al., 2006; UNEP and IOC-UNESCO, 2009; Diop et al., 2011). Coastal systems will experience impacts through sea level rise. They will also experience impacts through high sea levels combined with storm swells, for example as observed in Durban in March 2007, when a storm swell up to 14 m due to winds generated by a cyclone combined with a high astronomic tide at 2.2 m, leading to damages estimated at US$ 100 million (Mather and Stretch, 2012). Other climate change impacts (such as flooding of river deltas or an increased migration toward coastal towns due to increased drought induced by climate change (Rain et al., 2011) will also affect coastal zones. Some South African sea bird species have moved farther south over recent decades, but land use change may also have contributed to this migration (Hockey and Midgley, 2009; Hockey et al., 2011). However, it is considered that South African seabirds could be a valuable signal for climate change, particularly of the changes induced on prey species related to changes in physical oceanography, if we are able to separate the influences of climate parameters from other environmental ones (Crawford and Altwegg, 2009). Upwellings, including Eastern Boundary Upwelling Ecosystems (EUBEs) and Equatorial Upwelling Systems (EUSs) are the most biologically active systems in the Oceans (Box CC-UP). In addition to equatorial upwelling, the primary upwelling systems that affect Africa are the Benguela and Canary currents along the Atlantic coast (both EBUEs).The waters of the Benguela current have not shown warming over the period 1950-2009 (30.5.5.1.2), whereas most observations suggest that the Canary current has warmed since the early 1980s, and there is medium evidence and agreement that primary production in the Canary current has decreased over the past two decades (30.5.5.1.1). Changing temperatures in the Canary current has resulted in changes to important fisheries species (e.g., Mauritanian waters have become increasingly suitable for Sardinella aurita) (30.5.5.1.1). Upwellings are areas of naturally low pH and high CO2 concentrations, and, consequently, may be vulnerable to ocean acidification and its impacts (Box CC-OA, Box CC-UP, 30.5.5). Warming is projected to continue in the Canary current, and the synergies between this increase in water temperature and ocean acidification could influence a number of biological processes (30.5.5.2). Regarding the Benguela current upwelling, there is medium agreement despite limited evidence Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 that the Benguela system will experience changes in upwelling intensity as a result of climate change (30.5.5.1.2). There is considerable debate as to whether or not climate change will drive an intensification of upwelling (e.g. Bakun et al., 2010; Narayan et al., 2010) in all regions. Discussion of the various hypotheses for how climate change may affect coastal upwelling is presented in Box 30-1. Ocean acidification (OA) is the term used to describe the process whereby increased CO2 in the atmosphere, upon absorption, causes lowering of the pH of seawater (CC-OA). Projections indicate that severe impairment of reef accretion by organisms such as corals (Hoegh-Guldberg et al., 2007) and coralline algae (Kuffner et al., 2008) are substantial potential impacts of ocean acidification, and the combined effects of global warming and ocean acidification have been further demonstrated to lower both coral reef productivity (Anthony et al., 2008) and resilience (Anthony et al., 2011). These effects will have consequences for reef biodiversity, ecology, and ecosystem services (6.3.1, 6.3.2, 6.3.5, 6.4.1, 30.3.2, Box CC-CR). Coral vulnerability to heat anomalies is high in the Western Indian Ocean (30.5.6.1.2). Corals in the southwestern Indian Ocean (Comoros, Madagascar, Mauritius, Mayotte, Réunion and Rodrigues) appeared to be more resilient than those in eastern locations (30.5.6.1.2). Social adaptive capacity to cope with such change varies, and societal responses (such as closures to fishing) can have a positive impact on reef recovery, as observed in Tanzania (McClanahan et al., 2009). In Africa, fisheries mainly depend on either coral reefs (on the eastern coast) or coastal upwelling (on the western coast). These two ecosystems will be affected by climate change through ocean acidification, a rise in sea surface temperatures, and changes in upwelling (see Box CC-OA, Box CC-CR, Box CC- UP). [INSERT TABLE 22-3 Table 22-3: Examples of detected changes in species, natural ecosystems, and managed ecosystems in Africa that are both consistent with a climate change signal and published since the AR4. Confidence in detection of change is based on the length of study, and the type, amount and quality of data in relation to the natural variability in the particular species or system. Confidence in the role of climate being a major driver of the change is based on the extent to which the detected change is consistent with that expected under climate change, and to which other confounding or interacting non-climate factors have been considered and been found insufficient to explain the observed change.] 22.3.3. Water Resources Knowledge has advanced since the AR4 regarding current drivers of water resource abundance in Africa, and in understanding of potential future impacts on water resources from climate change and other drivers. However, inadequate observational data in Africa remains a systemic limitation with respect to fully estimating future freshwater availability (Neumann et al., 2007; Batisani, 2011). Detection of and attribution to climate change are difficult given that surface and groundwater hydrology are governed by multiple, interacting drivers and factors, such as land use change, water withdrawals, and natural climate variability (see also Chapter 3.2.1 and Box CC- WE). There is poor understanding in Africa of how climate change will affect water quality. This is an important knowledge gap. A growing body of literature generated since the AR4 suggests that climate change in Africa will have an overall modest effect on future water scarcity relative to other drivers, such as population growth, urbanization, agricultural growth, and land use change (high confidence) (Alcamo et al., 2007; Carter and Parker, 2009; MacDonald et al., 2009; Taylor et al., 2009; Calow and MacDonald, 2009; Abouabdillah et al., 2010; Beck and Bernauer, 2011; Droogers et al., 2012; Notter et al., 2012; Tshimanga and Hughes, 2012). However, broad-scale assumptions about drivers of future water shortages can mask significant sub-regional variability of climate impacts, particularly in water-stressed regions that are projected to become drier, such as northern Africa and parts of southern Africa. For example, rainfed agriculture in northern Africa is highly dependent on winter precipitation and would be negatively impacted if total precipitation and the frequency of wet days declines across North Africa as has been indicated in recent studies (Born et al., 2008; Driouech et al., 2010; Abouabdillah et al., 2010; García-Ruiz et al., 2011). Similarly, climate model predictions based on average rainfall years do not adequately capture interannual and Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 interdecadal variability that can positively or negatively influence surface water runoff (Beck and Bernauer, 2011; Notter et al., 2012; Wolski et al., 2012). Key challenges for estimating future water abundance in Africa lie in better understanding relationships between evapotranspiration, soil moisture, and land use change dynamics under varying temperature and precipitation projections (Goulden et al., 2009a) and to understand how compound risks such as heat waves and seasonal rainfall variability might interact in the future to impact water resources. Several studies from Africa point to a future decrease in water abundance due to a range of drivers and stresses, including climate change in Southern and northern Africa (medium confidence). For example, all countries within the Zambezi River Basin could contend with increasing water shortages (A2 scenario) although non-climate drivers (e.g., population and economic growth, expansion of irrigated agriculture, and water transfers) are expected to have a strong influence on future water availability in this basin (Beck and Bernauer, 2011). In Zimbabwe, climate change is estimated to increase water shortages for downstream users dependent on the Rozva dam (Ncube et al., 2011). Water shortages are also estimated for the Okavango Delta, from both climate change and increased water withdrawals for irrigation (Murray-Hudson et al., 2006; Milzow et al., 2010; Wolski et al., 2012), and the Breede River in South Africa (Steynor et al., 2009). For North Africa, Droogers et al. (2012) estimated that in 2050 climate change will account for 22% of future water shortages in the region while 78% of increased future water shortages can be attributed to socioeconomic factors. Abouabdillah et al. (2010) estimated that higher temperatures and declining rainfall (A2 and B1 scenarios) would reduce water resources in Tunisia. Reduced snowpack in the Atlas Mountains from a combination of warming and reduced precipitation, combined with more rapid springtime melting is expected to reduce supplies of seasonal meltwater for lowland areas of Morocco (García-Ruiz et al., 2011). In Eastern Africa, potential climate change impacts on the Nile Basin are of particular concern given the basin s geopolitical and socioeconomic importance. Reduced flows in the Blue Nile are estimated by late century due to a combination of climate change (higher temperatures and declining precipitation) and upstream water development for irrigation and hydropower (Elshamy et al., 2009; McCartney and Menker Girma, 2012). Beyene et al. (2010) estimated that streamflow in the Nile River will increase in the medium term (2010 2039) but will decline in the latter half of this century (A2 and B1 scenarios) as a result of both declining rainfall and increased evaporative demand, with subsequent diminution of water allocation for irrigated agriculture downstream from the High Aswan Dam. Kingston and Taylor (2010) reached a similar conclusion about an initial increase followed by a decline in surface water discharge in the Upper Nile Basin in Uganda. Seasonal runoff volumes in the Lake Tana Basin are estimated to decrease by the 2080s under the A2 and B2 scenarios (Abdo et al., 2009), while Taye et al. (2011) reported inconclusive findings as to changes in runoff in this basin. The Mara, Nyando, and Tana rivers in Eastern Africa, are projected to have increased flow in the second half of this century (Taye et al., 2011; Dessu and Melesse, 2012; Nakaegawa et al., 2012)). Estimating the influence of climate change on water resources in West Africa is limited by the significant climate model uncertainties with regards to the region s future precipitation. For example, Itiveh and Bigg (2008) estimate higher future rainfall in the Niger River Basin (A1, A2 and B1 scenarios), whereas Oguntunde and Abiodun (2013) report a strong seasonal component with reduced precipitation in the basin during the rainy season and increased precipitation during the dry season (A1B scenario). The Volta Basin is projected to experience a slight mean increase in precipitation (Kunstmann et al., 2008), and the Bani River Basin in Mali is estimated to experience substantial reductions in runoff (A2 scenario) due to reduced rainfall (Ruelland et al., 2012). The impact of climate change on total runoff in the Congo Basin is estimated to be minimal (A2 scenario) (Tshimanga and Hughes, 2012). Continental wide studies (e.g. De Wit and Stankiewicz, 2006) indicate that surface drainage in dry areas is more sensitive to, and will be more adversely affected by, reduced rainfall than would surface drainage in wetter areas that experience comparable rainfall reductions. The overall impact of climate change on groundwater resources in Africa is expected to be relatively small in comparison with impacts from non-climatic drivers such as population growth, urbanization, increased reliance on irrigation to meet food demand, and land use change (Calow and MacDonald, 2009; Carter and Parker, 2009; MacDonald et al., 2009; and Taylor et al., 2009). Climate change impacts on groundwater will vary across climatic zones. (See also Chapter 3.4.6). An analysis by MacDonald et al. (2009) indicated that changes in rainfall would not be expected to impact the recharge of deep aquifers in areas receiving below 200 mm rainfall per year, where recharge is negligible due to low rainfall. Groundwater recharge may also not be significantly affected by climate Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 change in areas that receive more than 500 mm per year, where sufficient recharge would remain even if rainfall diminished, assuming current groundwater extraction rates. By contrast, areas receiving between 200 to 500 mm per year, including the Sahel, the Horn of Africa, and southern Africa, may experience a decline in groundwater recharge with climate change to the extent that prolonged drought and other precipitation anomalies becomes more frequent with climate change, particularly in shallow aquifers, which respond more quickly to seasonal and yearly changes in rainfall than do deep aquifers (Barthel et al., 2009). Coastal aquifers are additionally vulnerable to climate change because of high rates of groundwater extraction, which leads to saltwater intrusion in aquifers, coupled with increased saltwater ingression resulting from sea level rise (Moustadraf et al., 2008; Bouchaou et al., 2008; Al-Gamal and Dodo, 2009; Kerrou et al., 2010). Some studies have shown additional impacts of sea level rise on aquifer salinization with salinity potentially reaching very high levels (Carneiro et al., 2010; Niang et al., 2010; Research Institute for Groundwater, 2011). Although these effects are expected to be localized, in some cases they will occur in densely populated areas (Niang et al., 2010). The profitability of irrigated agriculture in Morocco is expected to decline (under both B1 and A1B scenarios) due to increased pumping of groundwater and increased salinization risk for aquifers (Heidecke and Heckelei, 2010). The capacity of groundwater delivery systems to meet demand may take on increasing importance with climate change (Calow and MacDonald, 2009). For example, where groundwater pumping and delivery infrastructure is poor, and the number of point sources limited, prolonged pumping can lead to periodic drawdowns and increased failure of water delivery systems or increased saline intrusion (Moustadraf et al., 2008). To the extent that drought conditions become more prevalent in Africa with climate change, stress on groundwater delivery infrastructures will increase. Future development of groundwater resources to address direct and indirect impacts of climate change, population growth, industrialization, and expansion of irrigated agriculture, will require much more knowledge of groundwater resources and aquifer recharge potentials than currently exists in Africa. Observational data on groundwater resources in Africa are extremely limited and significant effort needs to be expended to assess groundwater recharge potential across the continent (Taylor et al., 2009). A preliminary analysis by MacDonald et al. (2012) indicates that total groundwater storage in Africa is 0.66 million km3, which is more than 100 times the annual renewable freshwater resources, and 20 times the freshwater stored in African lakes. However, borehole yields are variable and in many places water yields are relatively low. Detailed analysis of groundwater conditions for water resource planning would need to consider these constraints. 22.3.4. Agriculture and Food Security Africa s food production systems are among the world s most vulnerable because of extensive reliance on rainfed crop production, high intra- and inter-seasonal climate variability, recurrent droughts and floods that affect both crops and livestock, and persistent poverty that limits the capacity to adapt (Boko et al., 2007). In the near term, better managing risks associated with climate variability may help to build adaptive capacities for climate change (Washington et al., 2006; Cooper et al., 2008; Funk et al., 2008). However, agriculture in Africa will face significant challenges in adapting to climate changes projected to occur by mid-century, as negative effects of high temperatures become increasingly prominent under an A1B scenario (Battisti and Naylor, 2009; Burke et al., 2009a), thus increasing the likelihood of diminished yield potential of major crops in Africa (Schlenker and Lobell, 2010; Sultan et al., 2013). Changes in growing season length are possible, with a tendency towards reduced growing season length (Thornton et al., 2011), though with potential for some areas to experience longer growing seasons (Cook and Vizy, 2012). The composition of farming systems from mixed crop-livestock to more livestock dominated food production may occur as a result of reduced growing season length for annual crops and increases in the frequency and prevalence of failed seasons (Jones and Thornton, 2009; Thornton et al., 2010). Transition zones, where livestock keeping is projected to replace crop cultivation by 2050, include the West African Sahel and coastal and mid-altitude areas in eastern and southeastern Africa (Jones and Thornton, 2009), areas that currently support 35 million people and are chronically food insecure. Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.3.4.1. Crops Climate change is very likely to have an overall negative effect on yields of major cereal crops across Africa, with strong regional variability in the degree of yield reduction (see also Chapter 7.3.2.1) (Lobell et al., 2008; Liu et al., 2008; Walker and Schulze, 2008; Thornton et al., 2009a; Lobell et al., 2011; Roudier et al., 2011; Berg et al., 2013) (high confidence). One exception is in eastern Africa where maize production could benefit from warming at sites above roughly 1,700 m in elevation (A1FI scenario) (Thornton et al., 2009a), although the majority of current maize production occurs at lower elevations thereby implying a potential change in the distribution of maize cropping. Maize-based systems, particularly in southern Africa, are among the most vulnerable to climate change (Lobell et al., 2008). Estimated yield losses at mid-century range from 18% for southern Africa (Zinyengere et al., 2013) to 22% aggregated across SSA, with yield losses for South Africa and Zimbabwe in excess of 30% (Schlenker and Lobell, 2010). Simulations that combine all regions south of the Sahara suggest consistently negative effects of climate change on major cereal crops in Africa, ranging from 2% for sorghum to 35% for wheat by 2050 under an A2 scenario (Nelson et al., 2009). Studies in North Africa by Eid et al., 2007; Hegazy et al., 2008; Drine, 2011; Mougou et al., 2011 also indicate a high vulnerability of wheat production to projected warming trends. In West Africa, temperature increases above 2° C (relative to a 1961-1990 baseline) are estimated to counteract positive effects on millet and sorghum yields of increased precipitation (for B1, A1B and A2 scenarios) (Figure 22.5), with negative effects stronger in the savannah than in the Sahel, and with modern cereal varieties compared with traditional ones (Sultan et al., 2013). Several recent studies since the AR4 indicate that climate change will have variable impacts on non-cereal crops, with both production losses and gains possible (low confidence). Cassava yields in eastern Africa are estimated to moderately increase up to the 2030s assuming CO2 fertilization and under a range of low to high emissions scenarios (Liu et al., 2008), findings that were similar to Lobell et al. (2008). Suitability for growing cassava is estimated to increase with the greatest improvement in suitability in eastern and central Africa (A1B scenario) (Jarvis et al., 2012). However, Schlenker and Lobell (2010) estimated negative impacts from climate change on cassava at mid-century, although with impacts estimated to be less than those for cereal crops. Given cassava s hardiness to higher temperatures and sporadic rainfall relative to many cereal crops, it may provide a potential option for crop substitution of cereals as an adaptation response to climate change (Rosenthal and Ort, 2012; Jarvis et al., 2012). Bean yields in Eastern Africa are estimated to experience yield reductions by the 2030s under an intermediate emissions scenario (A1B) (Jarvis et al., 2012) and by the 2050s under low (B1) and high (A1FI) emissions scenarios (Thornton et al., 2011). For peanuts, some studies indicate a positive effect from climate change (A2 and B2 scenarios) (Tingem and Rivington, 2009) and others a negative one (Lobell et al., 2008; Schlenker and Lobell, 2010). Bambara groundnuts (Vigna subterranea) are estimated to benefit from moderate climate change (Tingem and Rivington, 2009) (A2 and B2 scenarios) although the effect could be highly variable across varieties (Berchie et al., 2012). Banana and plantain production could decline in West Africa and lowland areas of East Africa, whereas in highland areas of East Africa it could increase with temperature rise (Ramirez et al., 2011). Much more research is needed to better establish climate change impacts on these two crops. Suitable agro-climatic zones for growing economically important perennial crops are estimated to significantly diminish, largely due to the effects of rising temperatures (Läderach et al., 2010; Eitzinger et al., 2011a; Läderach et al., 2011a; Eitzinger et al., 2011b; Läderach et al., 2011b; Läderach et al., 2011c). Under an A2 scenario, by mid- century suitable agro-climatic zones that are currently classified as very good to good for perennial crops may become more marginal, and what are currently marginally suitable zones may become unsuitable; the constriction of crop suitability could be severe in some cases (Table 22-4). Movement of perennial crops to higher altitudes would serve to mitigate the loss of suitability at lower altitudes but this option is limited. Loss of productivity of high-value crops such as tea, coffee and cocoa would have detrimental impacts on export earnings. [INSERT TABLE 22-4 HERE Table 22-4. Projected changes in agro-climatic suitability for perennial crops in Africa by mid-century under an A2 scenario.] Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE 22-5 HERE Figure 22-5: The effect of rainfall and temperature changes on mean crop yield. Mean crop yield change (%) relative to the 1961 90 baseline for 7 temperatures (x-axis) and 5 rainfall (y-axis) scenarios. Results are shown as the average over the 35 stations across West Africa and the 6 cultivars of sorghum and millet. White triangles and circles are the projected anomalies computed by several CMIP3 GCMs and three IPCC emission scenarios (B1, A1B, A2) for 2071 90 and 2031 50, respectively. Projections from CMIP5 GCMs and three RCPs (4.5, 6.0 and 8.5) are represented by grey triangles and circles. Models and scenarios names are displayed in figure S2 (available atstacks.iop.org/ERL/8/014040/mmedia). Past observed climate anomalies from CRU data are also projected by computing 10-year averages (e.g. '1940' is for 1941 50). All mean yield changes are significant at a 5% level except boxes with a diagonal line. Source: Sultan et al., 2013.]. 22.3.4.2. Livestock Livestock systems in Africa face multiple stressors that can interact with climate change and variability to amplify the vulnerability of livestock-keeping communities. These stressors include rangeland degradation, increased variability in access to water, fragmentation of grazing areas, sedentarization, changes in land tenure from communal towards private ownership, in-migration of non-pastoralists into grazing areas, lack of opportunities to diversify livelihoods, conflict and political crisis, weak social safety nets, and insecure access to land, markets, and other resources (Solomon et al., 2007; Smucker and Wisner, 2008; Galvin, 2009; Thornton et al., 2009b; Dougill et al., 2010; Ifejika Speranza, 2010). (See also Chapter 7.3.2.4.) Loss of livestock under prolonged drought conditions is a critical risk given the extensive rangeland in Africa that is prone to drought. Regions that are projected to become drier with climate change, such as Northern and Southern Africa, are of particular concern (Solomon et al., 2007; Masike and Urich, 2008; Thornton et al., 2009b; Dougill et al., 2010; Freier et al., 2012; Schilling et al., 2012). Adequate provision of water for livestock production could become more difficult under climate change. For example, Masike and Urich (2009) estimated that the cost of supplying livestock water from boreholes in Botswana will increase by 23% by 2050 under an A2 scenario due to increased hours of groundwater pumping needed to meet livestock water demands under warmer and drier conditions. Although small in comparison to the water needed for feed production, drinking water provision for livestock is critical, and can have a strong impact on overall resource use efficiency in warm environments (Peden et al., 2009; van Breugel et al., 2010; Descheemaeker et al., 2010; Descheemaeker et al., 2011). Livestock production will be indirectly affected by water scarcity through its impact on crop production and subsequently the availability of crop residues for livestock feeding. Thornton et al. (2010) estimated that maize stover availability per head of cattle will decrease in several East African countries by 2050. The extent to which increased heat stress associated with climate change will affect livestock productivity has not been well established, particularly in the tropics and sub-tropics (Thornton et al., 2009b), although a few studies point to the possibility that keeping heat-tolerant livestock will become more prevalent in response to warming trends. For example, higher temperatures in lowland areas of Africa could result in reduced stocking of dairy cows in favor of cattle (Kabubo-Mariara, 2008), a shift from cattle to sheep and goats (Kabubo-Mariara, 2008; Seo and Mendelsohn, 2008), and decreasing reliance on poultry (Seo and Mendelsohn, 2008). Livestock keeping in highland areas of East Africa, which is currently cold-limited, would potentially benefit from increased temperatures (Thornton et al., 2010). Lunde and Lindtjrn (2013) challenge a finding in the AR4 that there is direct proportionality between range-fed livestock numbers and changes in annual precipitation in Africa. Their analysis indicates that this relationship may hold in dry environments but not in humid ones. 22.3.4.3. Agricultural Pests, Diseases, and Weeds Since the AR4, understanding of how climate change will potentially affect crop and livestock pests and diseases and agricultural weeds in Africa is beginning to emerge. Climate change in interaction with other environmental and production factors could intensify damage to crops from pests, weeds and diseases (7.3.2.3). Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Warming in highland regions of eastern Africa could lead to range expansion of crop pests into cold-limited areas (low confidence). For example, in highland arabica coffee-producing areas of eastern Africa, warming trends may result in the coffee berry borer (Hypothenemus hampei) becoming a serious threat in coffee-growing regions of Ethiopia, Kenya, Uganda, Rwanda, and Burundi (Jaramillo et al., 2011). Temperature increases in highland banana- producing areas of eastern Africa enhance the risk of altitudinal range expansion of the highly destructive burrowing nematode, Radopholus similis (Nicholls et al., 2008); however, no detailed studies have assessed this risk. Ramirez et al. (2011) estimated that increasing minimum temperatures by 2020 would expand the suitable range of Black Leaf Streak disease (Mycosphaerella fijiensis M.) of banana in Angola and Guinea. Climate change may also affect the distribution of economically important pests in lowland and dryland areas of Africa (low confidence). Under A2A and B2A for 2020, Cotter et al. (2012) estimated that changes in temperature, rainfall, and seasonality will result in more suitable habitats for Striga hermonthica in central Africa, whereas the Sahel region may become less suitable for this weed. Striga weed infestations are a major cause of cereal yield reduction in Sub-Saharan Africa. Climate change could also lead to an overall decrease in the suitable range of major cassava pests whitefly, cassava brown streak virus, cassava mosaic geminivirus, and cassava mealybug (Jarvis et al., 2012), although southeast Africa and Madagascar is estimated to experience increased suitability for cassava pests (Bellotti et al., 2012). In the case of livestock, Olwoch et al. (2008) estimated that the distribution of the main tick vector species (Rhipicephalus appendiculatus) of East Coast fever disease in cattle could be altered by a 2°C temperature increase over mean annual temperatures throughout the 1990s, and changes in mean precipitation resulting in the climatically suitable range of the tick shifting southward. However, a number of environmental and socio-economic factors (e.g., habitat destruction, land use and cover change, and host density) in addition to climatic ones influence tick distribution and need to be considered in assigning causality (Rogers and Randolph, 2006). 22.3.4.4. Fisheries Fisheries are an important source of food security in Africa. Capture fisheries (marine and inland) and aquaculture combined contribute over one-third of Africa s animal protein intake (Welcomme, 2011), while in some coastal countries fish contribute up to two-thirds of total animal protein intake (Allison et al., 2009). Demand for fish is projected to increase substantially in Africa over the next few decades (De Silva and Soto, 2009). To meet fish food demand by 2020, De Silva and Soto (2009) estimated that aquaculture production in Africa would have to increase nearly 500%. The vulnerability of national economies to climate change impacts on fisheries can be linked to exposure to the physical effects of climate change, the sensitivity of the country to impacts on fisheries, and adaptive capacity within the country (Allison et al., 2009). In an analysis of fisheries in 132 countries Allison et al. (2009) estimated that two- thirds of the most vulnerable countries were in Africa. Among these countries, the most vulnerable were Angola, DR Congo, Mauritania and Senegal, due to the importance of fisheries to the poor and the close link between climate variability and fisheries production. Coastal countries of West Africa will experience a significant negative impact from climate change. Lam et al. (2012) projected that by 2050 (under an A1B scenario) the annual landed value of fish for that region is estimated to decline by 21%, resulting in a nearly 50% decline in fisheries-related employment and a total annual loss of US$ 311 million to the region s economy. 22.3.4.5. Food Security Food security in Africa faces multiple threats stemming from entrenched poverty, environmental degradation, rapid urbanization, high population growth rates, and climate change and variability. The intertwined issues of markets and food security have emerged as an important issue in Africa and elsewhere in the developing world since the AR4. Price spikes for globally traded food commodities in 2007 2008 and food price volatility and higher overall food prices in subsequent years have undercut recent gains in food security across Africa (Brown et al., 2009; Hadley et al., 2011; Mason et al., 2011; Tawodzera, 2011; Alem and Söderbom, 2012; Levine, 2012). Among the Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 most affected groups are the urban poor, who typically allocate more than half of their income to food purchases (Cohen and Garrett, 2010; Crush and Frayne, 2010). The proportion of smallholder farmers that are net food buyers of staple grains exceeds 50% in Mozambique, Kenya and Ethiopia (Jayne et al., 2006), thus food security of rural producers is also sensitive to food spikes, particularly in the case of female-headed households, which generally have fewer assets than male-headed households (Kumar and Quisumbing, 2011). Although the recent spike in global food prices can be attributed to a convergence of several factors, the intensification of climate change impacts could become more important in the future in terms of exerting upward pressure on food prices of basic cereals (Nelson et al., 2009; Hertel et al., 2010), which would have serious implications for Africa s food security. As the recent wave of food price crises demonstrates, factors in other regions profoundly impact food security in Africa. Much more research is needed to understand better the potential interactions between climate change and other key drivers of food prices that act at national, regional, and global scales. (See also Chapter 7.2.2.) Africa is undergoing both rapid urbanization and subsequent transformation of its food systems to accommodate changes in food processing and marketing as well as in food consumption patterns. Considering the increasing reliance on purchased food in urban areas, approaches for addressing the impacts of climate change on food security will need to encompass a food systems approach (production as well as processing, transport, storage, and preparation) that moves food from production to consumption (Battersby, 2012). Weaknesses in the food system may be exacerbated by climate change in the region as high temperatures increase spoilage and the potential for increased flooding places food transportation infrastructure at higher risk of damage. In this respect, high post- harvest losses in Africa resulting in a large part from inadequate transport and storage infrastructure (Godfray et al., 2010; Parfitt et al., 2010) are an important concern. 22.3.5. Health 22.3.5.1. Introduction Africa currently experiences high burdens of health outcomes whose incidence and geographic range could be affected by changing temperature and precipitation patterns, including malnutrition, diarrheal diseases, and malaria and other vector-borne diseases, with most of the impact on women and children (WHO 2013a). In 2010, there were 451,000 to 813,000 deaths from malaria in Africa, continuing a slow decline since approximately 2004 (WHO, 2012). There are insufficient data series to assess trends in incidence in most affected countries in Africa. Parasite prevalence rates in children less than 5 years of age are highest in poorer populations and rural areas; factors increasing vulnerability include living in housing with little mosquito protection and limited access to health care facilities offering effective diagnostic testing and treatment. Of the 3.6 million annual childhood deaths in Africa, 11% are due to diarrheal diseases (Liu et al., 2012). Drivers of these and other climate-relevant health outcomes include inadequate human and financial resources, inadequate public health and health care systems, insufficient access to safe water and improved sanitation, food insecurity, and poor governance. Although progress has been made on improving safe water and sanitation coverage, sub-Saharan Africa still has the lowest coverage, highlighting high vulnerability to the health risks of climate change (UNICEF and WHO, 2008; UNICEF and WHO, 2012). Vulnerabilities also arise from policies and measures implemented in other sectors, including adaptation and mitigation options. Collaboration between sectors is essential. For example, the construction of the Akosombo dam in the 1960s to create Lake Volta in Ghana was associated with a subsequent increase in the prevalence of schistosomiasis (Scott et al., 1982). 22.3.5.2. Food- and Water-Borne Diseases Cholera is primarily associated with poor sanitation, poor governance, and poverty, with associations with weather and climate variability suggesting possible changes in incidence and geographic range with climate change (Rodó et al., 2002; Koelle et al., 2005; Olago et al., 2007; Murray et al., 2012). The frequency and duration of cholera outbreaks are associated with heavy rainfall in Ghana, Senegal, other coastal West African countries, and South Africa, with a possible association with the El Nino-Southern Oscillation (ENSO) (de Magny et al., 2007; Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mendelsohn and Dawson, 2008; de Magny et al., 2012). In Zanzibar, Tanzania, and Zambia, an increase in temperature or rainfall increases the number of cholera cases (Luque Fernández et al., 2009; Reyburn et al., 2011). The worst outbreak of cholera in recent African history occurred in Zimbabwe from August 2008 to June 2009. The epidemic was associated with the rainy season and caused more than 92,000 cases and 4,000 deaths. Contamination of water sources spread the disease (Mason, 2009). Poor governance, poor infrastructure, limited human resources, and underlying population susceptibility (high burden of malnutrition) contributed to the severity and extent of the outbreak (Murray et al., 2012). Other mechanisms for increases in cholera incidence have been described in Chapter 11 (11.5.2.1). As discussed above in section 22.2 there are projected increases in precipitation in areas in Africa for example West Africa where cholera is already endemic. This possibly will lead to more frequent cholera outbreaks in the sub-regions affected. However, further research is needed to quantify the climatic impacts. 22.3.5.3. Nutrition Malnutrition: Detailed spatial analyses of climate and health dynamics among children in Mali and Kenya suggest associations between livelihoods and measures of malnutrition, and between weather variables and stunting (Grace et al., 2012; Jankowska et al., 2012). Projections of climate and demographic change to 2025 for Mali (based on 2010-2039 climatology from the Famine Early Warning System Network FCLIM method), suggest approximately 250,000 children will suffer stunting, nearly 200,000 will be malnourished, and over 100,000 will become anemic, assuming constant morbidity levels; the authors conclude that climate change will cause a statistically significant proportion of stunted children (Jankowska et al., 2012). Using a process-driven approach, (Lloyd et al., 2011) projected future child malnutrition (as measured by severe stunting) in 2050 for four regions in sub-Saharan Africa, taking into consideration food and nonfood (socioeconomic) causes, and using regional scenario data based on the A2 scenario. Current baseline prevalence rates of severe stunting were 12-20%. Considering only future socioeconomic change, the prevalence of severe stunting in 2050 would be 7-17% (e.g. a net decline). However, including climate change, the prevalence of severe stunting would be 9-22%, or an increase of 31-55%in the relative percent of children severely stunted. Western sub- Saharan Africa was projected to experience a decline in severe stunting from 16% at present to 9% in 2050 when considering socioeconomic and climate change. Projected changes for central, south, and east sub-Saharan Africa are close to current prevalence rates, indicating climate change would counteract the beneficial consequences of socioeconomic development. Local economic activity and food accessibility can reduce the incidence of malnutrition (Funk et al., 2008; Rowhani et al., 2011). 22.3.5.4. Vector-Borne Diseases and Other Climate-Sensitive Health Outcomes A wide range of vector-borne diseases contribute to premature morbidity and mortality in Africa, including malaria, leishmaniasis, Rift Valley fever, as well as tick- and rodent-borne diseases. Malaria: Weather and climate are among the environmental, social, and economic determinants of the geographic range and incidence of malaria (Reiter 2008). The association between temperature and malaria varies regionally, (Chaves and Koenraadt, 2010; Paaijmans et al., 2010a; Alonso et al., 2011; Gilioli and Mariani, 2011). Malaria transmission peaks at 25°C and declines above 28°C (Lunde et al., 2013; Mordecai et al., 2013). Total precipitation, rainfall patterns, temperature variability, and the water temperature of breeding sites are expected to alter disease susceptibility (Bomblies and Eltahir, 2010; Paaijmans et al., 2010b; Afrane et al., 2012; Blanford et al., 2013; Lyons et al., 2013). ENSO events also may contribute to malaria epidemics (Mabaso et al., 2007; Ototo et al., 2011). The complexity of the malaria transmission cycle makes it difficult to determine whether the distribution of the pathogen and vector are already changing due to climate change. Other factors like the Indian Ocean Dipole have been proposed to affect malaria incidence (Hashizume et al., 2009, Chaves et al., 2012, Hashizume et al., 2012). Climate change is expected to affect the geographic range and incidence of malaria, particularly along the current edges of its distribution, with contractions and expansions, and increasing and decreasing incidence (Yé et al., 2007; Peterson, 2009; Parham and Michael, 2010; Paaijmans et al., 2010b; Alonso et al., 2011; Egbendewe-Mondzozo et Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 al., 2011; Chaves et al., 2012; Paaijmans et al., 2012; Parham et al., 2012; Ermert et al., 2012), depending on other drivers, such as public health interventions, factors influencing the geographic range and reproductive potential of malaria vectors, land use change (e.g., deforestation), and drug resistance, as well as the interactions of these drivers with weather and climate patterns (Chaves et al., 2008; Kelly-Hope et al., 2009; Paaijmans et al., 2009; Saugeon et al., 2009; Artzy-Randrup et al., 2010; Dondorp et al., 2010; Gething et al., 2010; Jackson et al., 2010; Kulkarni et al., 2010; Loha and Lindtjrn, 2010; Tonnang et al., 2010; Stern et al., 2011; Caminade et al., 2011; Omumbo et al., 2011; Afrane et al., 2012; Edlund et al., 2012; Githeko et al., 2012; Himeidan and Kweka, 2012; Jima et al., 2012; Lyons et al., 2012; Ermert et al., 2012; Stryker and Bomblies, 2012; Mordecai et al., 2013). Movement of the parasite into new regions is associated with epidemics with high morbidity and mortality. Because various Anopheles species are adapted to different climatic conditions, changing weather and climate patterns could affect species composition differentially, which could, in turn affect malaria transmission (Afrane et al., 2012; Lyons et al., 2013). Consensus is growing that highland areas, especially in East Africa, will experience increased malaria epidemics, with areas above 2,000 m, with temperatures currently too low to support malaria transmission, particularly affected (Pascual et al., 2006; Peterson, 2009; Gething et al., 2010; Lou and Zhao, 2010; Paaijmans et al., 2010a; Ermert et al., 2012). Reasons for different projections across models include use of different scenarios; use of global versus regional climate models (Ermert et al., 2012); the need for finer-scale and higher-resolution models of the sharp climate variations with altitude (Bouma et al., 2011); and the extent to malaria transmission and the drivers of its geographic range and incidence of malaria respond to and interact with climate change. Leishmaniasis: Directly or indirectly, climate change may increase the incidence and geographic range of leishmaniasis, a highly neglected disease that has recently become a significant health problem in northern Africa (Postigo, 2010), with a rising concern in western Africa because of co-infection with HIV (Kimutai et al., 2006). The epidemiology of the disease appears to be changing (Dondji, 2001; Yiougo et al., 2007; WHO, 2009; Postigo, 2010). During the 20th century, zoonotic cutaneous leishmaniasis emerged as an epidemic disease in Algeria, Morocco, and Tunisia, and is now endemic (Salah et al., 2007; Aoun et al., 2008; Rhajaoui, 2011; Toumi et al., 2012; Bounoua et al., 2013). Previously an urban disease in Algeria, leishmaniasis now has a peri-urban distribution linked to changes in the distribution of the rodent host and of the vector since the early 1990s (Aoun et al., 2008). Cutaneous leishmaniasis has expanded its range from its historical focus at Biskra, Algeria into the semi-arid steppe, with an associated upward trend in reported cases. In Morocco, sporadic cases of leishmania major (vector Phlebotomus papatasi) appeared early in the 20th century; since that time there have been occasional epidemics of up to 2,000 cases, interspersed with long periods with few or no cases (Rhajaoui, 2011). Outbreaks of zoonotic cutaneous leishmaniasis have become more frequent in Tunisia (where it emerged as an epidemic disease in 1991) (Salah et al., 2007; Toumi et al., 2012). The disease has since spread to adjacent areas in West Africa and East Africa (Dondji, 2001; Yiougo et al., 2007; WHO, 2009). Disease incidence is associated with rainfall and minimum temperature (Toumi et al., 2012; Bounoua et al., 2013). Relationships between decadal shifts over 1990-2009 in northwest Algeria and northeast Morocco in the number of cases and climate indicators suggested increased minimum temperatures created conditions suitable for endemicity (Bounoua et al., 2013). Environmental modifications, such as construction of dams, can change the temperature and humidity of the soil and thus affect vegetation that may result in changes in the composition and density of sandfly species and rodent vectors. More research however, is needed to quantity the climate related impacts because there are multiple underlying factors. Rift Vally fever (RVF): RVF epidemics in the Horn of Africa are associated with altered rainfall patterns. Additional climate variability and change could further increase its incidence and spread. Rift Valley fever is endemic in numerous African countries, with sporadic repeated epidemics. Epidemics in 2006 2007 in the Horn of Africa (Nguku et al., 2007; WHO, 2007; Adam et al., 2010; Andriamandimby et al., 2010; Hightower et al., 2012) and southern Africa were associated with heavy rainfall (Chevalier et al., 2011), strengthening earlier analyses by Anyamba et al. (2009) showing that RVF epizootics and epidemics are closely linked to the occurrence of the warm phase of ENSO and La Nina events (Linthicum et al., 1999; Anyamba et al., 2012) and elevated Indian Ocean temperatures. These conditions lead to heavy rainfall and flooding of habitats suitable for the production of the immature Aedes and Culex mosquitoes that serve as the primary RVF virus vectors in East Africa. Flooding of mosquito habitats also may introduce the virus into domestic animal populations. Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ticks and tick-borne diseases: Changing weather patterns could expand the distribution of ticks causing animal disease, particularly in East and South Africa. Ticks carry theileriosis (East Coast Fever), which causes anemia and skin damage that expose cattle to secondary infections. Habitat destruction, land use and cover change, and host density also affect tick distribution (Rogers and Randolph, 2006). Using a climate envelope and a species prediction model, Olwoch et al. (2007) projected that by 2020s, under the A2 scenario, East Africa and South Africa would be particularly vulnerable to climate-related changes in tick distributions and tick-borne diseases: more than 50% of the 30 Rhipicephalus species examined showed significant range expansion and shifts. More than 70% of this range expansion was found in tick species of economic importance. Schistosomiasis: Worldwide, approximately 243 million people required treatment for schistosomiasis in 2011, of which 90% lived in underdeveloped areas of Africa (WHO, 2013b). Water resource development, such as irrigation dams recommended for adaptation in agriculture, can amplify the risk of schistosomiasis (Huang and Manderson, 1992; Hunter et al., 1993; Jobin, 1999). Migration and sanitation play a significant role in the spread of schistosomiasis from rural areas to urban environments (Babiker et al., 1985; WHO, 2013b). Temperature and precipitation patterns may play a role in transmission (Odongo-Aginya et al., 2008; Mutuku et al., 2011; Huang et al., 2011). Projections for the period 2070-2099, under A2 and B2 emission scenarios, suggest that although the geographic areas suitable for transmission will increase with climate change, snail regions are expected to contract and/or move to cooler areas; these results highlight the importance of understanding how climate change could alter snail habitats when projecting future human schistosomiasis prevelance under different scenarios (Stensgaard et al. 2011). Meningococcal meningitis: There is a strong environmental relationship between the seasonal cycle of meningococcal meningitis and climate, including a relationship between the seasonal pattern of the Harmattan dusty winds and onset of disease. Transmission of meningitis occurs throughout Africa in the dry season and coincides with periods of very low humidity and wind-driven dusty conditions, ending with the onset of the rains (Molesworth et al., 2003). Research corroborates earlier hypothesized relationships between weather and meningitis (Yaka et al., 2008; Palmgren, 2009; Roberts, 2010; Dukiæ et al., 2012; Agier et al., 2013). In the northern region of Ghana, exposure to smoke from cooking fires increased the risk of contracting meningococcal meningitis (Hodgson et al., 2001). This increased risk suggests that exposure to elevated concentrations of air pollutants, such as carbon monoxide (CO) and particulate matter, may be linked to illness. More research is needed to clarify the possible impact of climate change on atmospheric concentrations of aerosols and particulates that can impact human health and any associations between meningitis and these aerosols and particles. The relationship between the environment and the location of the epidemics suggest connections between epidemics and regional climate variability (Molesworth et al., 2003; Sultan et al., 2005; Thomson et al., 2006) which may allow for early warning systems for predicting the location and onset of epidemics. Hantavirus: Novel hantaviruses with unknown pathogenic potential have been identified in some insectivores (shrews and a mole) in Africa (Klempa, 2009), with suggestions that weather and climate, among other drivers, could affect natural reservoirs and their geographic range, and thus alter species composition in ways that could be epidemiologically important (Klempa, 2009). Other health issues: Research into other health issues has begun. It has been noted that any increase in food insecurity due to climate change would be expected to further compromise the poor nutrition of people living with HIV/AIDs (Drimie and Gillespie, 2010). Laboratory studies suggest that the geographic range of the tsetse fly (Glossina species), the vector of human and animal trypanosomiasis in Africa, may be reduced with climate change (Terblanche et al., 2008). More studies are needed to clarify the role of climate change on HIV and other disease vectors. Heat waves and high ambient temperatures: Heat waves and heat-related health effects are only beginning to attract attention in Africa. High ambient temperatures are associated with increased mortality in Ghana, Burkina Faso, and Nairobi with associations varying by age, gender, and cause of death (Azongo et al., 2012; Diboulo et al., 2012; Egondi et al., 2012). Children are particularly at risk. Heat-related health effects also may be of concern in West and southern Africa (Dapi et al., 2010; Mathee et al., 2010). Chapter 11 (11.4.1) assesses the literature on the health impacts of heat waves and high ambient temperatures. Low ambient temperatures are associated with mortality in Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Nairobi and Tanzania (Egondi et al., 2012; Mrema et al., 2012). Chapter 11 discusses the relationship between heat and work capacity loss. This is an important issue for Africa because of the number of workers engaged in agriculture. Air quality: Climate change is anticipated to affect the sources of air pollutants as well as the ability of pollutants to be dispersed in the atmosphere (Denman et al., 2007). Assessments of the impacts of projected climate change on atmospheric concentrations of aerosols and particules that can adversely affect human health indicate that changes in surface temperature, land cover, and lightning may alter natural sources of ozone precursor gases and consequently ozone levels over Africa (Stevenson et al., 2005; Brasseur et al., 2006; Zeng et al., 2008). However, insufficient climate and emissions data for Africa prevent a more comprehensive assessment and further research is needed to better understand the implications of climate change on air quality in Africa. 22.3.6. Urbanization The urban population in Africa is projected to triple by2050, increasing by 0.8 billion (UN DESA, 2010). African countries are experiencing some of the world s highest urbanization rates (UN-Habitat, 2008). Many of Africa s evolving cities are unplanned and have been associated with growth of informal settlements, inadequate housing and basic services, and urban poverty (Yuen and Kumssa, 2011).7 [FOOTNOTE 7: However, community-driven upgrading may contribute to reducing the vulnerability of such informal areas (for more details see Chapter 8).] Climate change could affect the size and characteristics of rural and urban human settlements in Africa because the scale and type of rural-urban migration are partially driven by climate change (UN-Habitat and UNEP, 2010; Yuen and Kumssa, 2011). The majority of migration flows observed in response to environmental change are within country boundaries (Jäger et al., 2009; Tacoli, 2009). For large urban centers located on mega-deltas (e.g., Alexandria in Egypt in the Nile delta, and Benin City, Port Harcourt, and Aba in Nigeria in the Niger delta), urbanization through migration may also lead to increasing numbers of people vulnerable to coastal climate change impacts (Seto, 2011). Floods are exerting considerable impacts on cities and smaller urban centers in many African nations for example, heavy rains in East Africa in 2002 caused floods and mudslides, which forced tens of thousands to leave their homes in Rwanda, Kenya, Burundi, Tanzania and Uganda, and the very serious floods in Port Harcourt and Addis Ababa in 2006 (Douglas et al., 2008). Additionally sea level rise along coastal zones including coastal settlements could disrupt economic activities such as tourism and fisheries (Naidu et al., 2006; Kebede et al., 2012, Kebede and Nicholls, 2012). More than a quarter of Africa s population lives within 100 km of the coast and more than half of Africa s total population living in low- elevation coastal zones is urban, accounting for 11.5% of the total urban population of the continent (UN-Habitat, 2008). In eastern Africa, an assessment of the impact of coastal flooding due to sea level rise in Kenya found that, by 2030, 10,000 to 86,000 people would be affected, with associated economic costs ranging between US$ 7 million to US$ 58 million (SEI, 2009). Detailed assessments of damages arising from extreme events have also been made for some coastal cities, including Mombasa and Dar-es-Salaam. In Mombasa, by 2030 the population and assets at risk of 1 in 100-year return period extreme water levels is estimated to be between 170,700 to 266,300 inhabitants, while economic assets at risk are between US$ 0.68 billion and 1.06 billion (Kebede et al., 2012). In Dar-es-Salaam, the population and economic assets at risk of 1 in 100-year return period extreme water levels by 2030 range between 30,300 and 110,000 inhabitants and US$ 35.6 million to US$ 404.1 million (Kebede and Nicholls, 2012). For both city assessments, the breadth of these ranges encompasses three different population growth scenarios and four different sea-level rise scenarios (low (B1), medium (A1B), high (A1FI), and Rahmstorf (based on Rahmstorf, 2007)); these four sea-level rise scenarios were also the basis for the broader assessment of the coast of Kenya (SEI, 2009). The scale of the damages projected in the city-specific studies highlights the risks of extremes in the context of projected sea-level rise. Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In southern Africa, urban climate change risk assessments have been made at the regional scale (Theron and Rossouw, 2008) as well as at the city level for Durban, Cape Town, and the uMhlathuze local municipality. For these cities, risk assessments have focused on a broad range of sectors, including business and tourism; air quality, heath, and food security; infrastructure and services; biodiversity; and water resources (Naidu et al., 2006; Cartwright, 2008; Zitholele Consulting, 2009). Assessments for western Africa (Appeaning Addo et al., 2008; Niang et al., 2010) and northern Africa (Snoussi et al., 2009; World Bank, 2011) share similarities with those for eastern and southern Africa. For instance, it was suggested that by the end of the 21st century, about 23 %, 42 %, and 49 % of the total area of coastal governorates of the Nile Delta would be susceptible to inundation under the A1FI, Rahmstorf and Pfeffer scenarios of SLR. It was also suggested that a considerable proportion of these areas (ranging between 32% and 54 %) are currently either wetland or undeveloped areas (Hassaan and Abdrabo, 2013). Another study, assessing the economic impacts of sea level rise on the Nile Delta, suggested that losses in terms of housing and road would range between 1 and 2 billion EGP in 2030 and between 2 and 16 billion EGP in 2060 under the A1FI and B1 emissions scenarios as well as current sea level rise trends (Smith et al., 2013). African cities and towns represent highly vulnerable locations to the impacts of climate change and climate variability (Boko et al., 2007; Diagne, 2007; Dossou and Gléhouenou-Dossou, 2007; Douglas et al., 2008; Adelekan, 2010; Kithiia, 2011). Rapid rates of urbanization represent a burden on the economies of African urban areas, due to the massive investments needed to create job opportunities and provide infrastructure and services. Basic infrastructure services are not keeping up with urban growth, which has resulted in a decline in the coverage of many services, compared to 1990 levels (Banerjee et al., 2007). Squatter and poor areas typically lack provisions to reduce flood risks or to manage floods when they happen (Douglas et al., 2008). African small- and medium-sized cities have limited adaptive capacity to deal not only with future climate impacts but also with the current range of climate variability (Satterthwaite et al., 2009; UN-Habitat, 2011); for more details see Chapter 5 and Chapter 8). African cities, despite frequently having more services compared to rural areas (e.g., piped water, sanitation, schools and healthcare) that lead to human life spans above their respective national averages, show a shortfall in infrastructure due to low quality and short lifespan which may be of particular concern, when climate change impacts are taken into consideration (Satterthwaite et al., 2009). It is not possible, however, to climate-proof infrastructure that is not there (Satterthwaite et al., 2009). At the same time, hard infrastructural responses such as seawalls and channelized drainage lines are costly and can be maladaptive (Dossou and Gléhouenou-Dossou, 2007; Douglas et al., 2008; Kithiia and Lyth, 2011). High levels of vulnerability and low adaptive capacity results from structural factors, particularly local governments with poor capacities and resources (Kithiia, 2011). Weak local government creates and exacerbates problems including the lack of appropriate regulatory structures and mandates; poor or no planning; lack of or poor data; lack of disaster risk reduction strategies; poor servicing and infrastructure (particularly waste management and drainage); uncontrolled settlement of high-risk areas such as floodplains, wetlands, and coastlines; ecosystem degradation competing development priorities and timelines; and a lack of coordination among government agencies (AMCEN and UNEP, 2006; Diagne, 2007; Dossou and Gléhouenou-Dossou, 2007; Mukheibir and Ziervogel, 2007; Douglas et al., 2008; Roberts, 2008; Adelekan, 2010; Kithiia and Dowling, 2010; Kithiia, 2011). 22.4. Adaptation 22.4.1. Introduction Since 2007, Africa has gained experience in conceptualizing, planning and beginning to implement and support adaptation activities, from local to national levels and across a growing range of sectors (22.4.4, 22.4.5). However, across the continent, most of the adaptation to climate variability and change is reactive in response to short-term motivations, is occurring autonomously at the individual / household level, and lacks support from government stakeholders and policies (Vermuelen et al., 2008; Ziervogel et al., 2008; Berrang-Ford et al., 2011). A complex web Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 of interacting barriers to local-level adaptation, manifesting from national to local scales, both constrains and highlights potential limits to adaptation (22.4.6). 22.4.2. Adaptation Needs, Gaps, and Adaptive Capacity Africa s urgent adaptation needs stem from the continent s foremost sensitivity and vulnerability to climate change, together with its low levels of adaptive capacity (Ludi et al., 2012; 22.3). While overall adaptive capacity is considered low in Africa due to economic, demographic, health, education, infrastructure, governance and natural factors, levels vary within countries and across sub-regions, with some indication of higher adaptive capacity in North Africa and some other countries; individual or household level adaptive capacity depends, in addition to functional institutions and access to assets, on the ability of people to make informed decisions to respond to climatic and other changes (Vincent, 2007; Ludi et al., 2012). Inherent adaptation-related strengths in Africa include the continent s wealth in natural resources, well-developed social networks, and longstanding traditional mechanisms of managing variability through, for example, crop and livelihood diversification, migration and small-scale enterprises, all of which are underpinned by local or indigenous knowledge systems for sustainable resource management (Eyong, 2007; Nyong et al., 2007; UNFCCC, 2007; Cooper et al., 2008; Macchi et al., 2008; Nielsen, 2010; Castro et al., 2012;). However, it is uncertain to what extent these strategies will be capable of dealing with future changes, among them climate change and its interaction with other development processes (Leary et al., 2008b; Paavola, 2008; van Aalst et al., 2008; Conway, 2009; Jones, 2012, section 22.4.6). Since Africa is extensively exposed to a range of multiple stressors (22.3) that interact in complex ways ith longer term climate change, adaptation needs are broad, encompassing institutional, social, physical and infrastructure needs, ecosystem services and environmental needs, and financial and capacity needs. Making climate change information more reliable and accessible is one of the most pressing and cross-cutting adaptation needs, but providing information is insufficient to guarantee adaptation, which requires behavioural change (22.4.5.5, 22.4.6). As noted in the AR4 and emphasised in subsequent literature, monitoring networks in Africa are insufficient and characterised by sparse coverage and short and fragmented digitised records, which makes modelling difficult (Boko et al., 2007; Goulden et al., 2009b; Ziervogel and Zermoglio, 2009; Jalloh et al., 2011a). Adding to this is the shortage of relevant information and skills, in particular for downscaling climate models and using scenario outputs for development and adaptation planning, which is exacerbated by under- resourcing of Meteorological Agencies and a lack of in-country expertise on climate science; and the capacity of civil society and government organisations to access, interpret and use climate information for planning and decisionmaking (Ziervogel and Zermoglio, 2009; Brown et al., 2010; Ndegwa et al., 2010; Dinku et al., 2011; Jalloh et al., 2011a). Given its economic dependence on natural resources, most research on strengthening adaptive capacity in Africa is focused on agriculture, forestry or fisheries-based livelihoods (Collier et al., 2008; Berrang-Ford et al., 2011). The rural emphasis is now being expanded through a growing focus on requirements for enhancing peri-urban and urban adaptive capacity (Lwasa, 2010; Ricci, 2012). Many African countries have prioritised the following knowledge needs: vulnerability and impact assessments with greater continuity in countries; country-specific socio-economic scenarios and greater knowledge on costs and benefits of different adaptation measures; comprehensive programmes that promote adaptation through a more holistic development approach, including integrated programmes on desertification, water management and irrigation; promoting sustainable agricultural practices and the use of appropriate technologies and innovations to address shorter growing seasons, extreme temperatures, droughts, and floods; developing alternative sources of energy; and approaches to deal with water shortages, food security and loss of livelihoods (UNFCCC, 2007; Bryan et al., 2009; Eriksen and Silva, 2009; Chikozho, 2010; Gbetibouo et al., 2010b; Jalloh et al., 2011b; Sissoko et al., 2011; AAP, 2012). The literature, however, stresses the vast variety of contexts that shape adaptation and adaptive capacity - even when people are faced with the same climatic changes and livelihood stressors, responses vary greatly (Cooper et al., 2008; Vermuelen et al., 2008; Ziervogel et al., 2008; Gbetibouo, 2009; Westerhoff & Smit, 2009) Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Despite significant data and vulnerability assessment gaps, the literature highlights that delayed action on adaptation due to this would not be in the best interests of building resilience commensurate with the urgent needs (UNFCCC, 2007; Jobbins, 2011). See section 22.6.4 for a discussion of adaptation costs and climate finance. 22.4.3. Adaptation, Equity, and Sustainable Development Multiple uncertainties in the African context mean that successful adaptation will depend upon developing resilience in the face of uncertainty (high confidence) (Adger et al., 2011; Conway, 2011; Ludi et al., 2012). The limited ability of developmental strategies to counter current climate risks, in some cases due to significant implementation challenges related to complex cultural, political and insitutional factors, has led to an adaptation deficit, which reinforces the desirability for strong inter-linkages between adaptation and development, and for low-regrets adaptation strategies (see AR5 Glossary) that produce developmental co-benefits (high confidence) (Bauer and Scholz, 2010; Smith et al., 2011). Research has highlighted that no single adaptation strategy exists to meet the needs of all communities and contexts in Africa (high confidence) (22.4.4, 22.4.5). In recognition of the socioeconomic dimensions of vulnerability (Bauer and Scholz, 2010), the previous focus on technological solutions to directly address specific impacts is now evolving toward a broader view that highlights the importance of building resilience, through social, institutional, policy, knowledge, and informational approaches (ADF, 2010; Chambwera and Anderson, 2011), as well as on linking the diverse range of adaptation options to the multiple livelihood vulnerability risks faced by many people in Africa (Tschakert and Dietrich, 2010), and on taking into account local norms and practices in adaptation strategies (Nyong et al., 2007; Ifejika Speranza et al., 2010; section 22.4.5.4). Moreover, effective adaptation responses necessitate differentiated and targeted actions from the local to national levels, given the differentiated social impacts based on gender, age, disability, ethnicity, geographical location, livelihood, and migrant status (Tanner and Mitchell, 2008; IPCC, 2012). Additional attention to equity and social justice aspects in adaptation efforts in Africa, including the differential distribution of adaptation benefits and costs, would serve to enhance adaptive capacity (Burton et al., 2002; Brooks et al., 2005; Thomas and Twyman, 2005; Madzwamuse, 2010); nevertheless, some valuable experience has been gained recently on gender-equitable adaptation, human rights based-approaches, and involvement of vulnerable or marginalized groups such as indigenous peoples and children, aged and disabled people, internally displaced persons and refugees (ADF, 2010; UNICEF, 2010; Levine et al., 2011; UNICEF, 2011; Romero González et al., 2011; IDS, 2012; Tanner and Seballos, 2012) (Table 22-5). See also CC-GC on Gender and Climate Change. [INSERT TABLE 22-5 HERE Table 22-5: Cross-cutting approaches for equity and social justice in adaptation.] 22.4.4. Experiences in Building the Governance System for Adaptation, and Lessons Learned 22.4.4.1. Introduction Section 22.4.4 assesses progress made in developing policy, planning and institutional systems for climate adaptation at regional, national and sub-national levels in Africa, with some assessment of implementation. This includes an assessment of community-based adaptation, as an important local level response, and a consideration of adaptation decisionmaking and monitoring. 22.4.4.2. Regional and National Adaptation Planning and Implementation Regional policies and strategies for adaptation, as well as transboundary adaptation, are still in their infancy. Early examples include the Climate Change Strategies and Action Plans being developed by the Southern African Development Community and the Lake Victoria Basin Committee, as well as efforts being made by six highly Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 forested Congo basin countries to co-ordinate conservation and sustainable forest management of the Central African forest ecosystem, and obtain payments for ecosystem services (Harmeling et al., 2011; AfDB, 2012). At the national level, African countries have initiated comprehensive planning processes for adaptation by developing National Adaptation Programmes of Action (NAPAs), in the case of the Least Developed Countries, or National Climate Change Response Strategies (NCCRS); implementation is, however, lagging and integration with economic and development planning is limited but growing (high confidence). Prioritized adaptation measures in the NAPAs tend to focus narrowly on agriculture, food security, water resources, forestry, and disaster management; and onprojects, technical solutions, education and capacity development, with little integration with economic planning and poverty reduction processes (Madzwamuse, 2010; Mamouda, 2011; Pramova et al., 2012). Only a small percentage of the NAPA activities have been funded to date, although additional funding is in the pipeline (Prowse et al., 2009; Madzwamuse, 2010; Mamouda, 2011; Romero González et al., 2011). Subsequent to the NAPAs and early experience with the NCCRS, there is some evidence of evolution to a more integrated, multi-level and multi-sector approach to adaptation planning (medium confidence). Examples include Ethiopia s Programme of Adaptation to Climate Change, which includes sectoral, regional, national and local community levels (Hunde, 2012); Lesotho s co-ordinated policy framework involving all ministries and stakeholders (Corsi et al., 2012); and Mali s experience with a methodology for integrating adaptation into multiple sectors (Fröde et al., 2013). Cross-sectoral adaptation planning and risk management is occurring through mainstreaming initiatives like the twenty country Africa Adaptation Programme (AAP), initiated in 2008 (UNDP, 2009; Siegel, 2011). Examples of the more programmatic approach of national climate resilient development strategies include Rwanda s National Strategy on Climate Change and Low Carbon Development, under development in 2012, and the Pilot Programmes for Climate Resilience in Niger, Zambia and Mozambique (Climate Investment Fund, 2009). Inter-sectoral climate risk management approaches can be detected in integrated water resources management, integrated coastal zone management, disaster risk reduction, and land use planning initiatives (Boateng, 2006; Koch et al., 2007; Awuor et al., 2008; Cartwright et al., 2008; Kebede and Nicholls, 2011; Kebede et al., 2012), while in South Africa, climate change design principles have been incorporated into existing systematic biodiversity planning to guide land use planning (Petersen and Holness, 2011). The move to a more integrated approach to adaptation planning is occurring within efforts to construct enabling national policy environments for adaptation in many countries. Examples include Namibia s National Policy on Climate Change; Zambia s National Climate Change Response Strategy and Policy, and South Africa s National Climate Change Response Policy White Paper. Ten countries were developing new climate change laws or formal policies at the end of 2012, including the proposed National Coastal Adaptation Law in Gabon (Corsi et al., 2012). Despite this progress in mainstreaming climate risk in policy and planning, significant disconnects still exist at the national level, and implementation of a more integrated adaptation response remains tentative (high confidence) (Koch et al., 2007; Fankhauser and Schmidt-Traub, 2010; Madzwamuse, 2010; Oates et al., 2011; UNDP-UNEP Poverty-Environment Initiative, 2011a). Legislative and policy frameworks for adaptation remain fragmented, adaptation policy approaches seldom take into account realities in the political and institutional spheres, and national policies are often at odds with autonomous local adaptation strategies, which can act as a barrier to adaptation, especially where cultural, traditional and context-specific factors are ignored (Dube and Sekhwela, 2008; Patt and Schröter, 2008; Stringer et al., 2009; Bele et al. 2010; Hisali et al., 2011; Kalame et al. 2011; Naess et al., 2011; Lockwood, 2012; Sonwa et al. 2012; section 22.4.6). While climate resilience is starting to be mainstreamed into economic planning documents - for example, Zambia s Sixth National Development Plan 2011-2015, and the new Economic and Social Investment Plan in Niger (Corsi et al., 2012), measures to promote foreign direct investment and industrial competitiveness can undercut adaptive capacity of poor people (Madzwamuse, 2010), while poor business environments impede both foreign direct investment and adaptation (Collier et al., 2008). Stakeholders in climate-sensitive sectors for example, Botswana s tourism industry - have yet to develop and implement adaptation strategies (Saarinen et al., 2012). Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.4.4.3. Institutional Frameworks for Adaptation Global adaptation institutions, both within and outside of the UNFCCC, are critically important for Africa s ability to move forward on adaptation (14.2.3). Regional institutions focused on specific ecosystems rather than on political groupings, such as the Commission of Central African Forests (COMIFAC), present an opportunity to strengthen the institutional framework for adaptation. National frameworks include a number of institutions that cover all aspects of climate change: most countries have inter-ministerial coordinating bodies and inter-sectoral technical working groups, while an increasing number now have multi-stakeholder co-ordinating bodies (Harmeling et al, 2011) and are establishing national institutions to serve as conduits for climate finance (Gomez-Echeverri, 2010; Smith et al., 2011). Many studies in Africa show that under uncertain climatic futures, replacing hierarchical governance systems that operate within siloes with more adaptive, integrated, multi-level and flexible governance approaches, and with inclusive decisionmaking that can operate successfully across multiple scales or adaptive governance and co- management will enhance adaptive capacity and the effectivess of the adaptation response (Folke et al., 2005; Olsson et al., 2006; Koch et al., 2007; Berkes, 2009; Pahl-Wostl, 2009; Armitage and Plummer, 2010; Bunce et al., 2010a; Plummer, 2012). Despite some progress with developing the institutional framework for governing adaptation, there are significant problems with both transversal and vertical coordination, including institutional duplication with other inter-sectoral platforms, such for disaster risk reduction; while in fragile states, institutions for reducing climate risk and promoting adaptation may be extremely weak or almost non-existent (Hartmann and Sugulle, 2009; Sietz et al., 2011; Simane et al., 2012). Facilitating institutional linkages and co-ordinating responses across all boundaries of government, private sector and civil society would enhance adaptive capacity (Brown et al., 2010). Resolving well-documented institutional challenges of natural resource management, including lack of co- ordination, monitoring and enforcement, is a fundamental step towards more effective climate governance. For example, concerning groundwater, developing organizational frameworks and strengthening institutional capacities for more effectively assessing and managing groundwater resources over the long term are critically important (Nyenje and Batelaan, 2009; Braune and Xu, 2010). 22.4.4.4. Sub-National Adaptation Governance Since AR4, there has been additional effort on sub-national adaptation planning in African countries, but adaptation strategies at provincial and municipal levels are mostly still under development, with many local governments lacking the capacity and resources for the necessary decentralised adaptation response (high confidence). Provinces in some countries have developed policies and strategies on climate change: for example, Lagos State s 2012 Adaptation Strategy in Nigeria (BNRCC, 2012); mainstreaming adaptation into district development plans in Ghana; and communal climate resilience plans in Morocco (Corsi et al., 2012). Promising approaches include sub-national strategies that integrate adaptation and mitigation for low-carbon climate resilient development, as is being done in Delta State in Nigeria, and in other countries (UNDP, 2011a). In response to the identified institutional weaknesses, capacity development has been implemented in many cities and towns, including initiatives in Lagos, Nigeria, Durban and Cape Town in South Africa: notable examples include Maputo s specialized local government unit to implement climate change response, ecosystem-based adaptation and improved city wetlands; and participatory skills development in integrating community-based disaster risk reduction and climate adaptation into local development planning in Ethiopia (Madzwamuse, 2010; ACCRA, 2012; Castán Broto, et al., 2013). 22.4.4.5. Community-Based Adaptation and Local Institutions Since AR4, there has been progress in Africa in implementing and researching community-based adaptation (high confidence), with broad agreement that support to local-level adaptation is best achieved by starting with existing local adaptive capacity, and incorporating and building upon present coping strategies and norms, including indigenous practices (Dube and Sekhwela, 2007; Archer et al., 2008; Huq, 2011). Community-based adaptation (CBA) is community initiated, and/or draws upon community knowledge or resources refer to AR5 glossary. Some relevant initiatives include the Community-based Adaptation in Africa (CBAA) project, which implemented Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 community-level pilot projects in eight African countries (Sudan, Tanzania, Uganda, Zambia, Malawi, Kenya, Zimbabwe, South Africa) through a learning-by-doing approach; the Adaptation Learning Program, implemented in Ghana, Niger, Kenya and Mozambique (CARE International, 2012); and UNESCO Biosphere Reserves where good practices were developed in Ethiopia, Kenya, South Africa and Senegal (German Commission for UNESCO, 2011). See also section 22.4.5.6 on institutions for CBA. The literature includes a wide range of case studies detailing involvement of local communities in adaptation initiatives and projects facilitated by NGOs and researchers (for example, Leary et al., 2008a; CCAA, 2011; CARE International, 2012; Chishakwe et al., 2012); these and other initiatives have generated process-related lessons (22.4.5), with positive assessments of effectiveness in improving adaptive capacity of African communities, local organisations and researchers (Lafontaine et al., 2012). The key role for local institutions in enabling community resilience to climate change has been recognised, particularly with respect to natural resource dependent communities for example, the role of NGOs and CBOs in catalysing agricultural adaptation or in building resilience through enhanced forest governance and sustainable management of non-timber forest products; institutions managing access to and tenure of land and other natural resources, which are vital assets for the rural and peri-urban poor, are particularly crucial for enabling CBA and enhancing adaptive capacity in Africa (Bryan et al., 2009; Brown et al., 2010; Mogoi et al., 2010). Local studies and adaptation planning have revealed the following priorities for pro-poor adaptation: social protection, social services and safety nets; better water and land governance; action research to improve resilience of under-researched food crops of poor people; enhanced water storage and harvesting; better post-harvest services; strengthened civil society and greater involvement in planning; and more attention to urban and peri-urban areas heavily affected by migration of poor people (Moser and Satterthwaite, 2008; Urquhart, 2009; Bizikova et al., 2010). 22.4.4.6. Adaptation Decisionmaking and Monitoring Emerging patterns in Africa regarding adaptation decision-making, a critical component of adaptive capacity, include limited inclusive governance at the national level, with greater involvement in local initiatives of vulnerable and exposed people in assessing and choosing adaptation responses (high confidence). Civil society institutions and communities have to date played a limited role in formulation of national adaptation policies and strategies, highlighting the need for governments to widen the political space for citizens and institutions to participate in decision-making, for both effectiveness and to ensure rights are met (Madzwamuse, 2010; Castro et al., 2012). Building African leadership for climate change may assist with this (CCAA, 2011; Chandani, 2011; Corsi et al., 2012). A critical issue is how planning and decisionmaking for adaptation uses scientific evidence and projections, while also managing the uncertainties within the projections (Conway, 2011; Dodman and Carmin, 2011). A range of tools has been used in adaptation planning in Africa, including vulnerability assessment (22.4.5), risk assessment, cost-benefit analysis, cost-effectiveness, multi-criteria analysis, and participatory scenario planning (see for example Cartwright et al., 2008; Kemp-Benedict and Agyemang-Bonsu, 2008; Njie et al., 2008; Mather and Stretch, 2012), but further development and uptake of decision tools would facilitate enhanced decisionmaking. A related point is that monitoring and assessing adaptation is still relatively undeveloped in Africa, with national co- ordinating systems for collating data and synthesizing lessons not in place. Approaches for assessing adaptation action at local and regional levels have been developed (see for example Hahn et al., 2009; Gbetibouo et al., 2010a; Below et al., 2012), while there are positive examples of local monitoring of adaptation at the project level (see for example Archer et al., 2008; Below et al., 2012). Chapter 2 contains additional discussion of the foundations for decisionmaking on climate change matters. 22.4.5. Experiences with Adaptation Measures in Africa and Lessons Learned 22.4.5.1. Overview Section 22.4.5 provides a cross-cutting assessment of experience gained in Africa with a range of adaptation approaches, encompassing climate risk reduction measures; processes for participatory learning and knowledge Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 development and sharing; communication, education and training; ecosystem-based measures; and technological and infrastructural approaches; concluding with a discussion of maladaptation. Common priority sectors across countries for implementing adaptation measures since 2008 include agriculture, food security, forestry, energy, water, and education (Corsi et al., 2012), which reflects a broadening of focus since the AR4. While there has been little planning focus on regional adaptation (22.4.4.2, 22.4.4.3), the potential for this has been recognized (UNFCCC, 2007; Sonwa et al., 2009; Niang, 2012). Attention is increasing on identifying opportunities inherent in the continent s adaptation needs, as well as delineating key success factors for adaptation. A number of studies identify the opportunity inherent in implementing relatively low-cost and simple low-regrets adaptation measures that reduce people s vulnerability to current climate variability, have multiple developmental benefits, and are well-positioned to reduce vulnerability to longer-term climate change as well (UNFCCC, 2007; Conway and Schipper, 2011; see also Section 22.4.3). Responding to climate change provides an opportunity to enhance awareness that maintaining ecosystem functioning underpins human survival and development in a most fundamental way (Shackleton and Shackleton, 2012), and to motivate for new development trajectories (22.4.6). While it is difficult to assess adaptation success, given temporal and spatial scale issues, and local specificities, Osbahr et al. (2010) highlight the role of social networks and institutions, social resilience, and innovation as possible key success factors for adaptation in small- scale farming livelihoods in southern Africa. Kalame et al. (2008) note opportunities for enhancing adaptation through forest governance reforms to improve community access to forest resources, while Martens et al. (2009) emphasise the importance of soft path measures for adaptation strategies (see also section 22.4.5.6). The following discussion of adaptation approaches under discrete headings does not imply that these are mutually exclusive adaptation initiatives usually employ a range of approaches simultaneously, and indeed, the literature increasingly recognizes the importance of this for building resilience. 22.4.5.2. Climate Risk Reduction, Risk Transfer, and Livelihood Diversification Risk reduction strategies used in African countries to offset the impacts of natural hazards on individual households, communities, and the wider economy include early warning systems, emerging risk transfer schemes, social safety nets, disaster risk contingency funds and budgeting, livelihood diversification, and migration (World Bank, 2010; UNISDR, 2011). Disaster risk reduction (DRR) platforms are being built at national and local levels, with the synergies between DRR and adaptation to climate change being increasingly recognized in Africa (Westgate, 2010; UNISDR, 2011; Hunde, 2012); however, Conway and Schipper (2011) find that additional effort is needed for a longer-term vulnerability reduction perspective in disaster management institutions. Early warning systems (EWS) are gaining prominence as multiple stakeholders strengthen capabilities to assess and monitor risks and warn communities of a potential crisis, through regional systems such as the Permanent Inter- States Committee for Drought Control in the Sahel (CILSS) and the Famine Early Warning System Network (FEWS-NET), as well as national, local and community-based EWS on for example food and agriculture (Pantuliano and Wekesa, 2008; Sissoko et al., 2011; FAO, 2011). Some of the recent EWS emphasise a gendered approach, and may incorporate local knowledge systems used for making short-, medium-, and long-term decisions about farming and livestock-keeping, as in Kenya (UNDP, 2011b). The health sector has employed EWS used to predict disease for adaptation planning and implementation, such as the prediction of conditions expected to lead to an outbreak of Rift Valley fever in the Horn of Africa in 2006/2007 (Anyamba et al., 2010). Progress has been made in prediction of meningitis and in linking climate/weather variability and extremes to the disease (Thomson et al., 2006; Cuevas et al., 2007). Local projects often use participatory vulnerability assessment or screening to design adaptation strategies (van Vliet, 2010; GEF Evaluation Office, 2011; Hambira, 2011), but vulnerability assessment at the local government level is often lacking, and assessments to develop national adaptation plans and strategies have not always been Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 conducted in a participatory fashion (Madzwamuse, 2010). Kienberger (2012) details spatial modelling of social and economic vulnerability to floods at the district level in Búzi, Mozambique. Lessons from vulnerability analysis highlight that the highest exposure and risk do not always correlate with vulnerable ecosystems, socially marginalized groups, and areas with at-risk infrastructure, but may also lie in unexpected segments of the population (Moench, 2011). Community-level DRR initiatives include activities that link food security, household resilience, environmental conservation, asset creation, and infrastructure development objectives and co-benefits (Parry et al., 2009a; UNISDR, 2011; Frankenberger et al., 2012). Food security and nutrition-related safety nets and social protection mechanisms can mutually reinforce each other for DRR that promotes adaptation, as in Uganda s Karamoja Productive Assets Programme (Government of Uganda and WFP 2010; WFP, 2011). Initiatives in Kenya, South Africa, Swaziland and Tanzania have also sought to deploy local and traditional knowledge for the purposes of disaster preparedness and risk management (Mwaura, 2008; Galloway McLean, 2010). Haan et al. (2012) highlight the need for increased donor commitment to the resilience-building agenda within the framework of DRR, based on lessons from the 2011 famine in Somalia. Social protection8, a key element of the African Union social policy framework, is being increasingly used in Ethiopia, Rwanda, Malawi, Mozambique, South Africa, and other countries to buffer against shocks by building assets and increasing resilience of chronically and transiently poor households; in some cases this surpasses repeated relief interventions to address slower onset climate shocks, as in Ethiopia s Productive Safety Net Program (Brown et al., 2007; Heltberg et al. 2009). While social protection is helping with ex post and ex ante DRR and will be increasingly important for securing livelihoods should climate variability increase, less evidence exists for its effectiveness against the most extreme climatic shocks associated with higher emissions scenarios, which would require reducing dependence on climate-sensitive livelihood activities (Davies et al., 2009; Wiseman et al., 2009; Pelham et al., 2011; Béné et al., 2012). Social protection could further build adaptive capacity if based on improved understanding of the structural causes of poverty, including political and institutional dimensions (Brown et al., 2007; Davies et al., 2009; Levine et al., 2011). [FOOTNOTE 8: Social protection can include social transfers (cash or food), minimum standards such as for child labor, and social insurance.] Risk spreading mechanisms used in the African context include kinship networks; community funds; and disaster relief and insurance, which can provide financial security against extreme events such as droughts, floods, and tropical cyclones, and concurrently reduce poverty and enhance adaptive capacity9 (Leary et al., 2008a; Linnerooth- Bayer et al. 2009; Coe and Stern, 2011). Recent developments include the emergence of index-based insurance contracts (Box 22-1), which pay out not with the actual loss, but with a measurable event that could cause loss. [FOOTNOTE 9: Climate (or disaster) risk financing instruments include contingency funds; agricultural and property (private) insurance; sovereign insurance; reallocation of program expenditures; weather derivatives; and bonds.] _____ START BOX 22-1 HERE _____ Box 22-1. Experience with Index-Based Weather Insurance in Africa Malawi s initial experience of dealing with drought risk through index-based weather insurance directly to smallholders appears positive: 892 farmers purchased the insurance in the first trial period, which was bundled with a loan for groundnut production inputs (Hellmuth et al., 2009). In the next year, the pilot expanded, with the addition of maize, taking numbers up to 1,710 farmers and stimulating interest among banks, financiers, and supply chain participants such as processing and trading companies and input suppliers. A pilot insurance project in Ethiopia was designed to pay claims to the government based on a drought index that uses a time window between observed lack of rain and actual materialization of losses. This allows stakeholders to address threats to food security in ways that prevent the depletion of farmers productive assets, which reduces the future demand for humanitarian aid by enabling households to produce more food during subsequent seasons (Krishnamurty, 2011). Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Another key innovation in Ethiopia is the insurance for work program that allows cash-poor farmers to work for their insurance premiums by engaging in community-identified disaster risk reduction products, such as soil management and improved irrigation (WFP, 2011), which makes insurance affordable to the most marginalized and resource-poor sectors of society. _____ END BOX 22-1 HERE _____ The challenges associated with current risk reduction strategies include political and institutional challenges in translating early warning into early action (Bailey, 2013); communication challenges related to EWS: conveying useful information in local languages and communicating EWS in remote areas; national-level mistrust of locally collected data, which are perceived to be inflated to leverage more relief resources (Hellmuth et al., 2007; Pantuliano and Wekesa, 2008; Cartwright et al., 2008; FAO, 2011); the call for improved user-friendliness of early warning information, including at smaller spatial scales; the need for increased capacity in National Meteorological centres (22.4.2); and the need for better linkages between early warning, response, and prevention (Haan et al., 2012). Evidence is increasing that livelihood diversification, long used by African households to cope with climate shocks, can also assist with building resilience for longer term climate change by spreading risk. Over the past 20 years, households in the Sahel have reduced their vulnerability and increased their wealth through livelihood diversification, particularly when diversifying out of agriculture (Mertz et al., 2011). Households may employ a range of strategies, including on-farm diversification or specialization (Sissoko et al., 2011; Tacoli, 2011). Motsholapheko et al. (2011) show how livelihood diversification is used as an adaptation to flooding in the Okavango Delta, Botswana, and Badjeck et al. (2010) recommend private and public insurance schemes to help fishing communities rebuild after extreme events, and education and skills upgrading to enable broader choices when fishery activities can no longer be sustained. See Chapter 9 for a fuller discussion of the role of livelihood diversification in adaptation, particularly 9.3.3.1 and 9.3.5.2). Remittances are a longstanding and important means of reducing risk to climate variability and other household stressors, and of contributing to recovery from climatic shocks, as further discussed in Chapter 9 (9.3.3.3, 9.3.5.2). While livelihood diversification is an important adaptation strategy, it may replace formerly sustainable practices with livelihood activities that have negative environmental impacts (22.4.5.8). Rural finance and micro-credit can be enabling activities for adaptive response, which are also used by women for resilience-building activities (e.g., as documented in Sudan by Osman-Elasha et al., 2008). Credit and storage systems are instrumental in supporting families during the lean period, to prevent the sale of assets to buy food when market prices are higher (Romero González et al., 2011). Long seen as a fundamental process for most African families to incorporate choice into their risk profile and adapt to climate variability (Goldstone, 2002; Urdal, 2005; Reuveny, 2007; Fox and Hoelscher, 2010), there is evidence in some areas of the increased importance of migration (discussed in section 22.6.1, 8.2, 9.3.3.3, 12.4) and trade for livelihood strategies, as opposed to subsistence agriculture, as shown by Mertz et al. (2011) for the Sudano-Sahelian region of West Africa. 22.4.5.3. Adaptation as a Participatory Learning Process Since AR4, there has been more focus on the importance of flexible and iterative learning approaches for effective adaptation (medium evidence, high agreement). Due to the variety of intersecting social, environmental, and economic factors that affect societal adaptation, governments, communities, and individuals (Jones et al., 2010; Jones, 2012), adaptation is increasingly recognized as a complex process involving multiple linked steps at several scales, rather than a series of simple planned technical interventions (Moser and Ekstrom, 2010). Implementing adaptation as a participatory learning process enables people to adopt a proactive or anticipatory stance to avoid learning by shock (Tschakert and Dietrich, 2010). Iterative and experiential learning allows for flexible adaptation planning, appropriate considering the uncertainty inherent in climate projections that is compounded by other sources of flux affecting populations in Africa (Suarez Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 et al., 2008; Dodman and Carmin, 2011; Huq, 2011; Koelle and Annecke, 2011). Many studies have highlighted the utility of participatory action research, social and experiential learning, and creating enabling spaces for multi- stakeholder dialogue for managing uncertainty and unlocking the social and behavioral change required for adaptation (e.g., Tompkins and Adger, 2003; Ziervogel and Opere, 2010; Bizikova et al., 2010; Tschakert and Dietrich 2010; CCAA, 2011; Ebi et al., 2011; UNDP-UNEP Poverty-Environment Initiative, 2011b; Thorn, 2011; Faysse et al., 2012). Transdisciplinary approaches, which hold promise for enhancing linkages between sectors and thus reducing maladaptation are also starting to be adopted, as for example in the urban context (Evans, 2011). Learning approaches for adaptation may involve co-production of knowledge such as combining local and traditional knowledge with scientific knowledge (22.4.5.4). Adaptive co-management10 holds potential to develop capacity to deal with change (Watkiss et al., 2010; Plummer, 2012); the implications of strategic adaptive management for adaptation in aquatic protected areas in South Africa are being explored (Kingsford et al., 2011). [FOOTNOTE 10: Adaptive co-management is understood as a process by which institutional arrangements and ecological knowledge are tested and revised in a dynamic, ongoing, self-organized process of learning-by-doing (Folke et al., 2002).] Caveats and constraints to viewing adaptation as a participatory learning process include the time and resources required from both local actors and external facilitators, the challenges of multidisciplinary research, the politics of stakeholder participation and the effects of power imbalances, and the need to consider not only the consensus approach but also the role of conflicts (Aylett, 2010; Tschakert and Dietrich, 2010; Beardon and Newman, 2011; Jobbins, 2011; Shankland and Chambote, 2011). Learning throughout the adaptation process necessitates additional emphasis on ways of sharing experiences between communities and other stakeholders, both horizontally and vertically (22.4.5.4). Information and communication technologies, including mobile phones, radio, and the internet, can play a role in facilitating participatory learning processes and helping to overcome some of the challenges (Harvey et al., 2012). The increased emphasis on the importance of innovation for successful adaptation, in both rural and urban contexts, relates to interventions that employ innovative methods, as well as the innovation role of institutions (Tschakert and Dietrich, 2010; Dodman and Carmin, 2011; Rodima-Taylor, 2012; Scheffran et al., 2012). Scheffran et al. (2012) demonstrate how migrant social organizations in the western Sahel initiate innovations across regions by transferring technology and knowledge, as well as remittances and resources. While relevant, high-quality data is important as a basis for adaptation planning, innovative methods are being used to overcome data gaps, particularly local climatic data and analysis capability (Tschakert and Dietrich, 2010; GEF Evaluation Office, 2011). 22.4.5.4. Knowledge Development and Sharing Recent literature has confirmed the positive role of local and traditional knowledge in building resilience and adaptive capacity, and shaping responses to climatic variability and change in Africa (Nyong et al., 2007; Osbahr et al., 2007; Goulden et al., 2009b; Ifejika Speranza et al., 2010; Jalloh et al., 2011b; Newsham and Thomas, 2011). This is particularly so at the community scale, where there may be limited access to, quality of, or ability to use scientific information. The recent report on extreme events and disasters (IPCC, 2012) supports this view, finding high agreement and robust evidence of the positive impacts of integrating indigenous and scientific knowledge for adaptation. Concerns about the future adequacy of local knowledge to respond to climate impacts within the multi- stressor context include the decline in intergenerational transmission; a perceived decline in the reliability of local indicators for variability and change, as a result of socio-cultural, environmental, and climate changes (Hitchcock 2009; Jennings and Magrath 2009); and challenges of the emerging and anticipated climatic changes seeming to overrun indigenous knowledge and coping mechanisms of farmers (Berkes, 2009; Ifejika Speranza et al., 2010; Jalloh et al., 2011b; section 22.4.6). Based on analysis of the responses to the Sahel droughts during the 1970s and 1980s, Mortimore (2010) argues that local knowledge systems are more dynamic and robust than is often acknowledged. Linking indigenous and conventional climate observations can add value to climate change Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 adaptation within different local communities in Africa (Roncoli et al., 2002; Nyong et al., 2007; Chang'a et al., 2010; Guthiga and Newsham, 2011). Choosing specific adaptation actions that are informed by users perceptions and supported by accurate climate information, relevant to the scale where decisions are made, would be supportive of the largely autonomous adaptation taking place in Africa (Vogel and O Brien, 2006; Ziervogel et al., 2008; Bryan et al., 2009; Godfrey et al., 2010). Key problems regarding how science can inform decision making and policy are how best to match scientific information, for example about uncertainty of change, with decision needs; how to tailor information to different constituencies; and what criteria to use to assess whether or not information is legitimate to influence policy and decisionmaking (Vogel et al., 2007; Hirsch Hadorn et al., 2008). Institutional innovation is one solution: for example, Nigeria established the Science Committee on Climate Change to develop strategies to bridge the gap between increasing scientific knowledge and policy (Corsi et al., 2012). There is agreement that culture, or the shaping social norms, values, and rules including those related to ethnicity, class, gender, health, age, social status, cast, and hierarchy, is of crucial importance for adaptive capacity as a positive attribute but also as a barrier to successful local adaptation (22.4.6); further research is required in this field, not least because culture is highly heterogeneous within a society or locality (Adger et al., 2007, 2009; Ensor and Berger, 2009; Nielsen and Reenberg, 2010; Jones, 2012). Studies show that while it is important to further develop the evidence base for the effectiveness of traditional knowledge, integrating cultural components such as stories, myths, and oral history into initiatives to document local and traditional knowledge on adaptive or coping mechanisms is a key to better understanding how climate vulnerability and adaptation are framed and experienced (Urquhart, 2009; Beardon and Newman, 2011; Ford et al., 2012). Appropriate and equitable processes of participation and communication between scientists and local people have been found to prevent misuse or misappropriation of local and scientific knowledge (Nyong et al., 2007; Orlove et al., 2010; Crane, 2010). While multi-stakeholder platforms promote collaborative adaptation responses (CARE, 2012), adaptation initiatives in Africa lack comprehensive, institutionalised and proactive systems for knowledge sharing (GEF Evaluation Office, 2011; AAP, 2012). 22.4.5.5. Communication, Education, and Capacity Development Capacity development and awareness raising to enhance understanding of climate impacts and adaptation competencies and engender behavioural change have been undertaken through civil society-driven approaches or by institutions, such as regional and national research institutes, international and national programs and non- governmental organizations (UNFCCC, 2007; Reid et al., 2010; CCAA, 2011; START, 2011; Figueiredo and Perkins, 2012). Promising examples include youth ambassadors in Lesotho and civil society organizations in Tanzania (Corsi et al., 2012), and children as effective communicators and advocates for adaptation-related behavioral and policy change (22.4.3). Progress on inclusion of climate change into formal education is mixed, occurring within the relatively low priority given to environmental education in most countries (UNFCCC, 2007; Corsi et al., 2012; Mukute et al., 2012). Innovative methods used to communicate climate change include participatory video, photo stories, oral history videos, vernacular drama, radio, television and festivals, with an emphasis on the important rle of the media (Suarez et al., 2008; Harvey, 2011; Chikapa, 2012; Corsi et al., 2012). Better evidence-based communication processes will enhance awareness raising of the diverse range of stakeholders at all levels on the different aspects of climate change (Niang, 2007; Simane et al., 2012). A better understanding of the dimensions of the problem could be achieved by bringing together multiple users and producers of scientific and local knowledgein a trans-disciplinary process (Vogel et al., 2007; Hirsch Hadorn et al., 2008; Ziervogel et al., 2008; Koné et al., 2011) Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.4.5.6. Ecosystem Services, Biodiversity, and Natural Resource Management Africa s longstanding experiences with natural resource management, biodiversity use, and ecosystem-based responses such as afforestation, rangeland regeneration, catchment rehabilitation and community-based natural resource management (CBNRM) can be harnessed to develop effective and ecologically sustainable local adaptation strategies (high confidence). Relevant specific experiences include using mobile grazing to deal with both spatial and temporal rainfall variability in the Sahel (Djoudi et al., 2013); reducing the negative impacts of drought and floods on agricultural and livestock-based livelihoods through forest goods and services in Mali, Tanzania, and Zambia (Robledo et al., 2012); and ensuring food security and improved livelihoods for indigenous and local communities in West and Central Africa through the rich diversity of plant and animal genetic resources (Jalloh et al., 2011b). Natural resource management (NRM) practices that improve ecosystem resilience can serve as proactive, low regrets adaptation strategies for vulnerable livelihoods (high confidence). Two relevant widespread dual-benefit practices, developed to address desertification, are natural regeneration of local trees (Box 22-2) and water harvesting. Water harvesting practices11 have increased soil organic matter, improved soil structure, and increased agricultural yields at sites in Burkina Faso, Mali, Niger, and elsewhere, and are used by 60% of farmers in one area of Burkina Faso (Fatondji et al., 2009; Vohland and Barry, 2009; Barbier et al., 2009; Larwanou and Saadou, 2011). Although these and other practices serve as adaptations to climate change, revenue generation and other concerns may outweigh climate change as a motivating factor in their adoption (Mertz et al., 2009; Nielsen and Reenberg, 2010). While destocking of livestock during drought periods may also address desertification and adaptation, the lack of individual incentives and marketing mechanisms to destock and other cultural barriers inhibit their widespread adoption in the Sahel (Hein et al., 2009; Nielsen and Reenberg, 2010). Despite these provisos and other constraints (see for example Nelson and Agrawal, 2008; section 22.4.6 further highlights local-level institutional constraints), local stakeholder institutions for CBNRM do enable a more flexible response to changing climatic conditions; CBNRM is also a vehicle for improving links between ecosystem services and poverty reduction, to enable sustainable adaptation approaches (Shackleton et al., 2010; Chishakwe et al., 2012; Girot et al., 2012). Based on lessons learned in Botswana, Malawi, Mozambique, Namibia, Tanzania, Zambia and Zimbabwe, Chishakwe et al. (2012) point out the synergies between CBNRM and adaptation at the community level, notwithstanding institutional and other constraints experienced with CBNRM. [FOOTNOTE 11: Water harvesting refers to a collection of traditional practices in which farmers use small planting pits, half-moon berms, rock bunds along contours, and other structures to capture runoff from episodic rain events (Kandji et al., 2006).] _____ START BOX 22-2 HERE _____ Box 22-2. African Success Story: Integrating Trees into Annual Cropping Systems Recent success stories from smallholder systems in Africa illustrate the potential for transforming degraded agricultural landscapes into more productive, sustainable and resilient systems by integrating trees into annual cropping systems. For example, in Zambia and Malawi, an integrated strategy for replenishing soil fertility on degraded lands, which combines planting of nitrogen-fixing Faidherbia trees with small doses of mineral fertilizers, has consistently more than doubled yields of maize leading to increased food security and greater income generation (Garrity et al., 2010). In the Sahel, natural regeneration, or the traditional selection and protection of small trees to maturity by farmers and herders has, perhaps for centuries, produced extensive parks of Acacia albida (winter thorn) in Senegal (Lericollais, 1989), Adansonia digitata (baobab) in West and southern Africa (Sanchez et al., 2011), and Butyrospermum parkii (Shea butter) in Burkina Faso (Gijsbers et al., 1994). Recent natural regeneration efforts have increased tree density and species richness at locations in Burkina Faso (Raebild et al., 2012) and Niger (Larwanou and Saadou, 2011), though adoption and success is somewhat dependent on soil type (Haglund et al., 2011; Larwanou and Saadou, 2011). In southern Niger, farmer-managed natural regeneration of Faidherbia albida and other field trees, which began in earnest in the late 1980s, has led to large-scale increase in tree cover across 4.8 million ha, and to decreased sensitivity to drought of the production systems, compared to other regions in Niger (Reij et al., 2009; Tougiani et al., 2009; Sendzimir et al., 2011). Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ END BOX 22-2 HERE _____ Differentiation in the literature is growing between hard path and soft path approaches to adaptation (Sovacool, 2011; Kundzewicz, 2011), with soft path , low-regrets approaches, such as using intact wetlands for flood risk management, often the first line of defence for poor people in Africa; as contrasted with hard path approaches like embankments and dams for flood control (McCully, 2007; Kundzewicz, 2011). Intact ecosystem services and biodiversity are recognized as critical components of successful human adaptation to climate change that may be more effective and incur lower costs than hard or engineered solutions (Abramovitz et al., 2002; Petersen and Holness, 2011; UNDP-UNEP Poverty-Environment Initiative, 2011a; Girot et al, 2012; Pramova et al., 2012; Roberts et al., 2012; Box 22-2). This provides a compelling reason for linking biodiversity, developmental, and social goals, as taken up, for example, in Djibouti s NAPA project on mangrove restoration to reduce salt water intrusion and coastal production losses due to climate hazards (Pramova et al., 2012). The emerging global concept of ecosystem-based adaptation (EbA) provides a system-oriented approach for Africa s longstanding local NRM practices. Despite the evidence from studies cited in this section, scaling-up to prioritize ecosystem responses and EbA in plans and policy has been slow; a broad understanding that EbA is an integral component of the developmental agenda, rather than a competing green agenda, would promote this process. Adaptive environmental governance represents one of the future challenges for the implementation of EbA strategies in Africa, together with sustainable use of resources, secure access to meet needs under climate change, and strong local institutions to enable this (Robledo et al., 2012). Ecosystem-based adaptation could be an important approach to consider for the globally significant Congo Basin forests, particularly given the predominance of REDD+ approaches for this region that risk neglecting adaptation responses, or may result in maladaptation (Somorin et al., 2012; Sonwa et al., 2012; sections 22.4.5.8, 22.6.2). Ecosystem-based approaches are further discussed in WGII Chapter 4, and Box CC-EA. 22.4.5.7. Technological and Infrastructural Adaptation Responses Since AR4, experience has been gained on technological and infrastructural adaptation in agricultural and water management responses, for climate-proofing infrastructure, and for improved food storage and management to reduce post-harvest losses; this has been increasingly in conjunction with soft measures. There is increased evidence that farmers are changing their production practices in response to increased food security risks linked to climate change and variability, through both technical and behavioural means. Examples include planting cereal crop varieties that are better suited to shorter and more variable growing seasons (Akullo et al., 2007; Thomas et al., 2007; Yesuf et al., 2008; Yaro, 2010; Laube et al., 2012), constructing bunds to more effectively capture rainwater and reduce soil erosion (Nyssen et al., 2007; Thomas et al., 2007; Reij et al., 2009), reduced tillage practices and crop residue management to more effectively bridge dry spells (Ngigi et al., 2006; Marongwe et al., 2011), and adjusting planting dates to match shifts in the timing of rainfall (Abou-Hadid, 2006; Vincent et al., 2011b). Conservation agriculture has good potential to both bolster food production and enable better management of climate risks (high confidence) (Verchot et al., 2007; Thomas, 2008; Syampungani et al., 2010; Thierfelder and Wall, 2010; Kassam et al., 2012). Such practices, which include conservation/zero tillage, soil incorporation of crop residues and green manures, building of stone bunds, agroforestry, and afforestation/reforestation of croplands, reduce runoff and protect soils from erosion, increase rainwater capture and soil water-holding capacity, replenish soil fertility, and increase carbon storage in agricultural landscapes. Conservation agriculture systems have potential to lower the costs of tillage and weed control with subsequent increase in net returns, as found in Malawi by Ngwira et al. (2012). Expansion of irrigation in sub-Saharan Africa holds significant potential for spurring agricultural growth while also better managing water deficiency risks associated with climate change (Dillon, 2011; You et al., 2011). Embedding irrigation expansion within systems-level planning that considers the multi-stressor context in which irrigation Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 expansion is occurring can help to ensure that efforts to promote irrigation can be sustained and do not instead generate a new set of hurdles for producers or engender conflict (van de Giesen et al., 2010; Burney and Naylor, 2012; Laube et al., 2012). Suitable approaches to expand irrigation in Africa include using low-pressure drip irrigation technologies and construction of small reservoirs, both of which can help to foster diversification toward irrigated high-value horticultural crops (Karlberg et al., 2007; Woltering et al., 2011; Biazin et al., 2012). If drought risk increases and rainfall patterns change, adaptation in agricultural water management would be enhanced through a strategic approach that encompasses overall water use efficiency for both rainfed and irrigated production (Wei et al., 2009), embeds irrigation expansion efforts within a larger rural development context that includes increased access to agricultural inputs and markets (You et al., 2011; Burney and Naylor, 2012), and that involves an integrated suite of options (e.g., plant breeding and improved pest and disease and soil fertility management, and in situ rainwater harvesting) to increase water productivity (Passioura, 2006; Biazin et al., 2012). Experience has been gained since the AR4 on adaptation of infrastructure (transportation, buildings, food storage, coastal), with evidence that this can sometimes be achieved at low cost, and additional implementation of soft measures such as building codes and zone planning (UNFCCC, 2007; Halsnaes and and Trarup, 2009; Urquhart, 2009; UN-Habitat and UNEP, 2010; AfDB, 2011; Mosha, 2011; Siegel, 2011; Corsi et al., 2012). Examples of adaptation actions for road and transportation infrastructure include submersible roads in Madagascar and building dikes to avoid flooding in Djibouti (UNFCCC, 2007; Urquhart, 2009). Infrastructural climate change impact assessments and enhanced construction and infrastructural standards - such as raising foundations of buildings, strengthening roads, and increasing storm water drainage capacity - are steps to safeguard buildings in vulnerable locations or with inadequate construction (UN-Habitat and UNEP, 2010; Mosha, 2011; Corsi et al., 2012). Mainstreaming adaptation into infrastructure development can be achieved at low cost, as has been shown for flood- prone roads in Mozambique (Halsnaes and and Trarup, 2009). Integrating climate change considerations into infrastructure at the design stage is preferable from a cost and feasibility perspective than trying to retrofit infrastructure (Chigwada, 2005; Siegel, 2011). Softer measures, such as building codes and zone planning are being implemented and are needed to complement and/or provide strategic guidance for hard infrastructural climate proofing, for example, the adoption of cyclone-resistant standards for public buildings in Madagascar (AfDB, 2011). Research in South Africa has recognized that the best option for adaptation in the coastal zone is not to combat coastal erosion in the long term, but rather to allow progression of the natural processes (Naidu et al., 2006; Zitholele Consulting, 2009). Reducing post-harvest losses through improved food storage, food preservation, greater access to processing facilities, and improved systems of transportation to markets are important means to enhance food security (Brown et al., 2009; Godfray et al., 2010; Codjoe and Owusu 2011). Low cost farm-level storage options, such as metal silos (Tefera et al., 2011), and triple-sealed plastic bags (Baoua et al., 2012) are effective for reducing post-harvest losses from pests and pathogens. Better storage allows farmers greater flexibility in when they sell their grain, with related income benefits (Brown et al., 2009), and reduces post-harvest infection of grain by aflatoxins, which is widespread in Africa and increases with drought stress and high humidity during storage (Cotty and Jaime-Garcia, 2007; Shephard, 2008). 22.4.5.8. Maladaptation Risks The literature increasingly highlights the need, when designing development or adaptation research, policies and initiatives, to adopt a longer-term view and to consider the multi-stressor context in which people live, in order to avoid maladaptation, or outcomes that may serve short-term goals but come with future costs to society (refer to Glossary). The short-term nature of policy and other interventions, especially if they favor economic growth and modernization over resilience and human security, may themselves act as stressors or allow people to only react to short-term climate variability (Bryan et al., 2009; Brooks et al., 2009; Bunce et al., 2010a; Levine et al, 2011). The political context can also undermine autonomous adaptation and lead to maladaptation; for instance, Smucker and Wisner (2008) found that political and economic changes in Kenya meant that farmers could no longer use traditional strategies for coping with climatic shocks and stressors, with the poorest increasingly having to resort to coping strategies that undermined their long-term livelihood security, also known as erosive coping, such as more intensive grazing of livestock and shorter crop rotations (van der Geest and Dietz, 2004). In a case from the Simiyu Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 wetlands in Tanzania, Hamisi et al. (2012) find that coping and reactive adaptation strategies may lead to maladaptation for instance, through negative impacts on natural vegetation because of increased intensity of farming in wetter parts of the floodplain, where farmers have moved to exploit the higher soil water content. Some diversification strategies, such as charcoal production and artisanal mining, may increase risk through promoting ecological change and the loss of ecosystem services to fall back on (Paavola, 2008; Adger et al, 2011; Shackleton and Shackleton, 2012). Studies also highlight risks that traditional adaptive pastoralism systems may be replaced by maladaptive activities. For example, charcoal production has become a major source of income for 70% of poor and middle-income pastoralists in some areas of Somaliland, with resultant deforestation (Hartmann and Sugulle, 2009). Another example of maladaptation provided in the literature is the potential long-term hydro-dependency risks and threats to ecosystem health and community resilience as a result of increased dam building in Africa, which may be underpinned by policies of multi-lateral donors (Avery, 2012; Beilfuss, 2012; Jones et al., 2012). While increased rainwater storage will assist with buffering dry periods, and hydropower can play a key role in ending energy poverty, it is important that this is designed to promote environmental and social sustainability; that costs and benefits are equitably shared; and that water storage and energy generation infrastructure is itself climate proofed. Additional substantive review of such international development projects would assist in assuring that these do not result in maladaptation. See WGII Chapter 4 for a discussion of the unwanted consequences of building more and larger impoundments and increased water abstraction on terrestrial and freshwater ecosystems; health aspects of this are noted in sections 22.3.5.1 and 22.3.5.4. See section 22.6.2 on avoiding undesirable trade-offs between REDD+ approaches and adaptation that have the potential to result in significant maladaptation. 22.4.6. Barriers and Limits to Adaptation in Africa A complex web of interacting barriers to local-level adaptation exists that manifests from national to local scales to constrain adaptation, which includes institutional, political, social, cultural, biophysical, cognitive and behavioral, and gender-related (high confidence). While relatively few studies from Africa have focused specifically on barriers and limits to adaptation, perceived and experienced constraints distilled from the literature encompass the resources needed for adaptation, the factors influencing adaptive capacity, the reasons for not employing particular adaptive strategies or not responding to climate change signals, and the reasons why some groups or individuals adapt but not others (Roncoli et al., 2010; Bryan et al., 2011; Nyanga et al., 2011; Ludi et al., 2012). At the local level, institutional barriers hamper adaptation through elite capture and corruption; poor survival of institutions without social roots; and lack of attention to the institutional requirements of new technological interventions (Ludi et al., 2012). Tenure security over land and vital assets is widely accepted as being crucial for enabling people to make longer-term and forward-looking decisions in the face of uncertainty, such as changing farming practices, farming systems, or even transforming livelihoods altogether (Bryan et al., 2009; Brown et al., 2010; Romero González et al., 2011). In addition to unclear land tenure, legislation forbidding ecosystem use is one of the issues strengthening underlying conflicts over resources in Africa; resolving this would enable ecosystems to contribute to adaptation beyond short-term coping (Robledo et al., 2012). There is also evidence that innovation may be suppressed if the dominant culture disapproves of departure from the normal way of doing things (Ludi et al., 2012; Jones 2012). Characteristics such as wealth, gender, ethnicity, religion, class, caste, or profession can act as social barriers for some to adapt successfully or acquire the required adaptive capacities (Ziervogel et al., 2008; Godfrey et al. 2010; Jones and Boyd, 2011). Based on field research conducted in the Borana area of southern Ethiopia, Debsu (2012) highlights the complex way in which external interventions may affect local and indigenous institutions by strengthening some coping and adaptive mechanisms and weakening others. Restrictive institutions can block attempts to enhance local adaptive capacity by maintaining structural inequities related to gender and ethnic minorities (Jones, 2012). Constraints faced by women, often through customs and legal barriers, include limited access to land and natural resources; lack of credit and input in decisionmaking, limited ability to take financial risk, lack of confidence, limited access to information and new ideas, and under-valuation of women s opinions Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (McFerson, 2010; Peach Brown, 2011, Djoudi and Brockhaus 2011; Jones 2012; Ludi et al., 2012; Goh, 2012; Codjoe et al. 2012). Few small-scale farmers across Africa are able to adapt to climatic changes, while others are restricted by a suite of overlapping barriers (high agreement, robust evidence). Constraints identified in Kenya, South Africa, Ethiopia, Malawi, Mozambique, Zimbabwe, Zambia and Ghana included poverty and a lack of cash or credit (financial barriers); limited access to water and land, poor soil quality, land fragmentation, poor roads, and pests and diseases (biophysical and infrastructural barriers); lack of access to inputs, shortage of labor, poor quality of seed and inputs attributed to a lack of quality controls by government and corrupt business practices by traders, insecure tenure, and poor market access (institutional, technological, and political barriers); and finally a lack of information on agroforestry/afforestation, different crop varieties, climate change predictions and weather, and adaptation strategies (informational barriers) (Bryan et al. 2009, 2011; Barbier et al., 2009; Clover and Eriksen, 2009; Deressa et al., 2009; Roncoli et al., 2010; Mandleni and Anim, 2011; Nhemachena and Hassan, 2011; Nyanga et al.,2011; Vincent et al., 2011a). Recognition is increasing that understanding psychological factors such as mindsets and risk perceptions is crucial for supporting adaptation (Grothmann and Patt, 2005; Patt and Schröter, 2008; Jones, 2012). Cognitive barriers to adaptation include alternative explanations of extreme events and weather such as religion (God s will), the ancestors, and witchcraft, or seeing these changes as out of people s own control (Byran et al., 2009; Roncoli et al., 2010; Mandleni and Anim, 2011; Artur and Hilhorst, 2012; Jones, 2012; Mubaya et al. 2012). Climate uncertainty, high levels of variability, lack of access to appropriate real-time and future climate information, and poor predictive capacity at a local scale are commonly cited barriers to adaptation from the individual to national level (Repetto, 2008; Dinku et al., 2011; Jones, 2012; Mather and Stretch, 2012). Despite the cultural and psychological barriers noted above, several studies have shown that farmers with access to climate information are more predisposed to adjust their behaviour in response to perceived climate changes (Mubaya et al., 2012). At a policy level, studies have detected political, institutional and discursive barriers to adaptation. Adaptation options in southern Africa have been blocked by political and institutional inefficiencies, lack of prioritization of climate change, and the dominance of other discourses, such as the mitigation discourse in South Africa and short- term disaster-focused views of climate variability (Madzwamuse, 2010; Bele et al., 2011; Berrang-Ford et al., 2011; Conway and Schipper, 2011; Kalame et al., 2011; Chevallier, 2012; Toteng, 2012; Leck et al., 2012). Lack of local participation in policy formulation, the neglect of social and cultural context, and the inadvertent undermining of local coping and adaptive strategies have also been identified by several commentators as barriers to appropriate national policies and frameworks that would support local-level adaptation (e.g., Brockhaus and Djoudi, 2008; Bele et al., 2011; Chevallier, 2012). Many of these constraints to adaptation are well entrenched and will be far from easy to overcome; some may act as limits to adaptation for particular social groups (high confidence). Biophysical barriers to adaptation in the arid areas could present as limits for more vulnerable groups if current climate change trends continue (Leary et al., 2008b; Sallu et al., 2010; Roncoli et al., 2010). Traditional and autonomous adaptation strategies, particularly in the drylands, have been constrained by social-ecological change and drivers such as population growth, land privatization, land degradation, widespread poverty, HIV/AIDS, poorly conceived policies and modernization, obstacles to mobility and use of indigenous knowledge, as well as erosion of traditional knowledge, to the extent that it is difficult or no longer possible to respond to climate variability and risk in ways that people did in the past (Leary et al., 2008b; Dabi et al., 2008; Paavola, 2008; Smucker and Wisner, 2008; Clover and Eriksen, 2009; Conway, 2009; UNCCD et al., 2009; Bunce et al., 2010b; Quinn et al., 2011; Jones, 2012; section 22.4.5.4). As a result of these multiple stressors working together, the number of response options has decreased and traditional coping strategies are no longer sufficient (Dube and Sekhwela, 2008). Studies have shown that most autonomous adaptation usually involves minor adjustments to current practices (e.g. changes in planting decisions); there are simply too many barriers to implementing substantial changes that require investment (e.g., agroforestry and irrigation) (Bryan et al., 2011). Such adaptation strategies would be enhanced through government and private Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 sector/NGO support, without which many poor groups in Africa may face real limits to adaptation (Vincent et al., 2011a; Jones, 2012). These findings highlight the benefits of transformational change in situations where high levels of vulnerability and low adaptive capacity detract from the possibility for systems to adapt sustainably. This is in agreement with the Special Report on Extreme Events, which additionally found high agreement and robust evidence for the importance of a spectrum of actions ranging from incremental steps to transformational changes in order to reduce climate risks (IPCC, 2012). In support of such solutions, Moench (2011) has called for distilling common principles for building adaptive capacity at different stages, and adaptive management and learning are seen as critical approaches for facilitating transformation (Section 22.4.5.3; IPCC, 2012). Chapter 16 provides further discussion on how encountering limits to adaptation may trigger transformational change, which can be a means of adapting to hard limits. 22.5. Key Risks for Africa Table 22-6 highlights key risks for Africa (see also Table 19-4 and CC-KR), as identified through assessment of the literature and expert judgement of the author team, with supporting evaluation of evidence and agreement in the sections of this chapter, as referenced in the caption. [INSERT TABLE 22-6 HERE Table 22-6: Key risks from climate change and the potential for risk reduction through mitigation and adaptation in Africa. Key risks are identified based on assessment of the literature and expert judgments made by authors of the various WGII AR5 chapters, with supporting evaluation of evidence and agreement in the referenced chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels are presented for the near-term era of committed climate change (here, for 2030-2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the longer-term era of climate options (here, for 2080-2100), for global mean temperature increase of 2°C and 4°C above preindustrial levels. For each timeframe, risk levels are estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols.] As indicated in Table 22-6, seven of the nine key regional risks are assessed for the present as being either medium or high under current adaptation levels, reflecting both the severity of multiple relevant stressors and Africa s existing adaptation deficit. This is the case for risks relating to shifts in biome distribution (22.3.2.1), degradation of coral reefs (22.3.2.3), reduced crop productivity (22.3.4.1), adverse effects on livestock (22.3.4.2), vector- and water-borne diseases (22.3.5.2, 22.3.5.4), under nutrition (22.3.5.3), and migration (22.6.1.2). This assessment indicates that allowing current emissions levels to result in a +4°C world (above pre-industrial levels) by the 2080- 2100 period would have negative impacts on Africa s food security, as even under high adaptation levels, risks of reduced crop productivity and adverse effects on livestock are assessed as remaining very high. Moreover, our assessment is that even if high levels of adaptation were achieved, risks of stress on water resources (22.3.3), degradation of coral reefs (22.3.2.3), and the destructive effects of sea level rise and extreme weather events (22.3.6) would remain high. However, even under a lower emissions scenario leading to a long-term 2°C warming, all nine key regional risks are assessed as remaining high or very high under current levels of adaptation. The assessment indicates that even under high adaptation, residual impacts in a 2°C world would be significant, with only risk associated with migration rated as being capable of reduction to low under high levels of adaptation. High adaptation would be enabled by concerted effort and substantial funding; even if this is realized, no risk is assessed as being capable of reduction to below medium status. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 22.6. Emerging Issues 22.6.1. Human Security Although the significance of human security cannot be overestimated, the evidence of the impact of climate change on human security in Africa is disputable (see Chapters 12 and 19). Adverse climate events potentially impact all aspects of human security, either directly or indirectly (on mapping climate security vulnerability in Africa see Busby et al., 2013). Food security, water stress, land use and, health security, violent conflict, changing migration patterns, and human settlements are interrelating issues of discussions between climate change and human security. Violent conflict and migration are discussed below (for further details see Chapter 12). 22.6.1.1 Violent Conflict While there seems to be consensus that the environment is only one of several interconnected causes of conflict and is rarely considered to be the most decisive factor (Kolmannskog, 2010), it remains disputed whether, and if so, how, the changing climate directly increases the risk of violent conflict in Africa (for more details see also Chapters 12 and 19, in particular 12.5.1; 19.4; Gleditsch, 2012). However, views are emerging that there is a positive relationship between increases in temperature and increases in human conflict (Hsiang et al., 2013). Some of the factors which may increase the risk of violent conflict, such as low per capita incomes, economic contraction and inconsistent state institutions are sensitive to climate change (12.5.1). For the African Sahel States it has been argued that the propensity for communal conflict across ethnic groups within Africa is influenced by political and economic vulnerability to climate change (Raleigh, 2010). Evidence on the question of whether, and if so to what extent, climate change and variability increases the risk of civil war in Africa is contested (Burke et al., 2009b; Buhaug, 2010; Devitt and Tol, 2012). It has been suggested that due to the depletion of natural resources in Africa as a result of overexploitation and the impact of climate change on environmental degradation, competition for scarce resources could increase and lead to violent conflict (Kumssa and Jones, 2010). For East Africa it has been suggested that increased levels of malnutrition are related to armed conflicts (Rowhani et al., 2011). There is some agreement that rainfall variation has an inconsistent relationship to conflict: both higher and lower anomalous rainfall is associated with increased communal conflict levels; although dry conditions have a lesser effect (Raleigh and Kniveton, 2012; Hendrix and Salehyan 2012; Theisen 2012). 22.6.1.2 Migration Human migration has social, political, demographic, economic and environmental drivers, which may operate independently or in combination (for more in-depth discussions see Chapters 12.4 and 19.4.2.1; Perch-Nielsen et al., 2008; Piguet, 2010; Foresight, 2011; Piguet et al., 2011; Black et al., 2011a; Van der Geest, 2011). Many of these drivers are climate sensitive (Black et al., 2011c; 12.4.1.). People migrate either temporarily or permanently, within their country or across borders (12.4.1.2; Figure 12-1; Table 12-3; Warner et al., 2010; Kälin and Schrepfer, 2012). The evidence base in the field of migration in Africa is both varied and patchy. Evidence suggests that migration is a strategy to adapt to climate change (12.4.2). Mobility is indeed a strategy (not a reaction) to high levels of climatic variation that is characteristic of Africa (Tacoli, 2011) and the specifics of the response are determined by the economic context of the specific communities. Besides low-lying islands and coastal and deltaic regions in general, sub-Saharan Africa is one of the regions that would particularly be affected by environmentally induced migration (Gemenne, 2011a). Case studies from Somalia and Burundi emphasize the interaction of climate change, disaster, conflict, displacement, and migration (Kolmannskog, 2010). In Ghana for example, an African country with few conflicts caused by political, ethnic, or religious tensions, and thus with migration drivers more likely related to economic and environmental motivators (Tschakert and Tutu, 2010), some different types of migration flows are considered to have different sensitivity to climate change (Black et al., 2011a). The floods of the Zambezi River in Mozambique in 2008 have displaced 90,000 people, and it has been observed that along the Zambezi River Valley, with approximately 1 million people Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 living in the flood-affected areas, temporary mass displacement is taking on permanent characteristics (Jäger et al., 2009; Warner et al., 2010). Different assessments of future trends have recently produced contradictory conclusions (e.g., UN-OCAH and IDMC, 2009; Naude, 2010; ADB, 2011; Tacoli, 2011, IDMC, 2011). One approach in assessing future migration potentials, with considerable relevance to the African context, focused on capturing the net effect of environmental change on aggregate migration through analysis of both its interactions with other migration drivers and the role of migration within adaptation strategies, rather than identifying specific groups as potential environmental migrants (Foresight, 2011). Even if Africa s population doubles by 2050 to 2 billion (Lutz and K.C., 2010) and the potential for displacement rises as a consequence of the impact of extreme weather events, recent analyses (Foresight, 2011; Black et al., 2011b) show that the picture for future migration is much more complex than previous assessments of a rise in climate induced migration suggest, and relates to the intersection of multiple drivers with rates of global growth, levels of governance, and climate change. The empirical base for major migration consequences is weak (Liller and Van den Broeck, 2011; Black et al., 2011a; Gemenne, 2011b) and non-existent for international migration patterns (Marchiori et al., 2011). Even across the same type of extreme weather event, the responses can vary (Findlay, 2011; Gray, 2011 for Kenya and Uganda; Raleigh, 2011 for the African Sahel States)). 22.6.2. Integrated Adaptation / Mitigation Approaches Relevant experience gained in Africa since AR4 in implementing integrated adaptation mitigation responses within a pro-poor orientation that leverages developmental benefits encompasses some participation of farmers and local communities in carbon offset systems, increasing the use of relevant technologies such as agroforestry and farmer- assisted tree regeneration (22.4.5.6), and emerging Green Economy policy responses. The recognition that adaptation and mitigation are complementary elements of the global response to climate change, and not trade-offs, is gaining traction in Africa (Goklany, 2007; Nyong et al., 2007; UNCCD et al., 2009; Woodfine, 2009; Jalloh et al., 2011b; Milder et al., 2011). While the suitability of on- and off-farm techniques for an integrated adaptation-mitigation response depends on local physical conditions as well as political and institutional factors, sustainable land management techniques are particularly beneficial for an integrated response in Africa; these include agroforestry, including through farmer- managed natural regeneration; and conservation agriculture (Woodfine, 2009; Milder et al., 2011; Mutonyi and Fungo, 2011; section 22.4.5.6; Box 22-2). An emerging area is multiple-benefit initiatives that aim to reduce poverty, promote adaptation through restoring local ecosystems, and deliver benefits from carbon markets. Brown et al., (2011) note the example of a community-based project in Humbo, Ethiopia, which is facilitating adaptation and generating temporary certified emissions reductions under the Clean Development Mechanism, by restoration of degraded native forests (2,728 ha) through farmer-managed natural regeneration. The key role of local communities in carbon offset systems through community forestry entails land use flexibility (Purdon, 2010), but can be constrained by the lack of supportive policy environments for example, for conservation agriculture (Milder et al., 2011). The literature highlights the desirability of responding to climate change through integrated adaptation mitigation approaches, including through spatial planning, in the implementation of REDD+ in Africa, especially given the significant contribution to food security and livelihoods of forest systems (Bwango et al., 2000; Guariguata et al., 2008; Nkem et al., 2007; Nasi et al., 2008; Biesbroek et al., 2009; Somorin et al., 2012). However, forests are mainly used for reactive coping and not anticipatory adaptation; studies show that governments favour mitigation while local communities prioritise adaptation (Fisher et al., 2010; Somorin et al., 2012). Flexible REDD+ models that include agriculture and adaptation hold promise for generating co-benefits for poverty reduction, given food security and adaptation priorities, and help to avoid trade-offs between REDD+ implementation and adaptive capacities of communities, ecosystems, and nations (Nkem et al., 2008; Thomson et al., 2010; CIFOR, 2011; Richard et al., 2011; Wertz-Kanounnikoff et al., 2011). Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Integrated adaptation mitigation responses are being considered within the context of the emerging Green Economy discussions. African leaders agreed in 2011 to develop an African Green Growth Strategy, to build a shared vision for promoting sustainable low-carbon growth through a linked adaptation mitigation approach, with adaptation seen as an urgent priority. A national example is the launch of Ethiopia s Climate Resilient Green Economy Facility in 2012 (Corsi et al., 2012). 22.6.3. Biofuels and Land Use The potential for first-generation biofuel production in Africa, derived from bioethanol from starch sources and biodiesel production from oilseeds, is significant given the continent s extensive arable lands, labor availability, and favorable climate for biofuel crop production (Amigun et al., 2011; Arndt et al., 2011; Hanff et al., 2011). While biofuel production has positive energy security and economic growth implications, the prospect of wide-scale biofuel production in Africa carries with it significant risks related to about environmental and social sustainability. Among the concerns are competition for land and water between fuel and food crops, adverse impacts of biofuels on biodiversity and the environment, contractual and regulatory obligations that expose farmers to legal risks, changes in land tenure security, and reduced livelihood opportunities for women, pastoralists and migrant farmers who depend on access to the land resource base (Unruh, 2008; Amigun et al., 2011; German et al., 2011; Schoneveld et al., 2011). More research is needed to understand fully the socioeconomic and environmental tradeoffs associated with biofuel production in Africa. One critical knowledge gap concerns the effect of biofuel production, particularly large-scale schemes, on land use change and subsequent food and livelihood security. For example, the conversion of marginal lands to biofuel crop production would impact the ability of users of these lands (pastoralists and in some cases women who are allocated marginal land for food and medicinal production) to participate in land use and food production decisions (Amigun et al., 2011; Schoneveld et al., 2011). In addition, biofuel production could potentially lead to the extension of agriculture into forested areas, either directly through conversion of fallow vegetation or the opening of mature woodland, or indirectly through use of these lands to offset food crop displacement (German et al., 2011). Such land use conversion would result in biofuel production reducing terrestrial carbon storage potential (Vang Rasmussen et al., 2012a; Vang Rasmussen et al., 2012b). Better agronomic characterization of biofuel crops is another key knowledge gap. For example, little information exists with respect to the agronomic characteristics of the oilseed crop Jatropha (Jatropha curcas) under conditions of intensive cultivation across differing growing environments, despite the fact that Jatropha has been widely touted as an appropriate feedstock for biofuel production in Africa because of its ability to grow in a wide range of climates and soils. Oilseed yields of Jatropha can be highly variable, and even basic information about yield potential and water and fertilizer requirements for producing economically significant oilseed yields is scanty (Achten et al., 2008; Peters and Thielmann, 2008; Hanff et al., 2011). Such knowledge would not only provide a basis for better crop management but would also help to gain better estimates of the extent of water consumption for biofuel production in the context of non-biofuel water-use needs across landscapes. Assessments of Jatropha s potential as an invasive species and its potential allelopathic effects on native vegetation are also needed, in light of the fact that some countries have designated Jatropha as an invasive species (Achten et al., 2008). 22.6.4. Climate Finance and Management Recent analyses emphasise the significant financial resources and technological support needed to both address Africa s current adaptation deficit and to protect rural and urban livelihoods, societies and economies from climate change impacts at different local scales, with estimates of adaptation costs between US$20-30 billion per annum over the next couple of decades, up to US$60 billion per annum by 2030 (for example see figure 22.6), although these figures are likely to be under-estimates, as studies upon which these estimates are based do not always include the costs of overcoming Africa s current adaptation deficit, may be run for one scenario at a time, and do not factor Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 in a range of uncertainties in the planning environment (Parry et al, 2009b; Fankhauser and Schmidt-Traub, 2010; Watkiss et al, 2010; AfDB, 2011; Dodman and Carmin, 2011; LDC Expert Group, 2011; Smith et al, 2011). Damages related to climate change may affect economic growth and the ability to trade (Lecocq and Shalizi, 2007; Ruppel and Ruppel-Schlichting, 2012). Costs of adaptation and negative economic impacts of climate change have been referred to in sections 22.3.4.4 and 22.3.6; Warner et al. (2012) have highlighted the residual impacts of climate change that would occur after adaptation, for case studies in Kenya and The Gambia. The following examples are illustrative of the move to discuss financial implications in the literature. Scenarios for Tanzania, where agriculture accounts for about half of gross production and employs about 80% of the labor force (Thurlow and Wobst, 2003), project that changes in the mean and extremes of climate variables, could increase poverty vulnerability (Ahmed et al., 2011). Scenarios for Namibia based on a computable general equilibrium model project that annual losses to the economy ascribed to the impacts of climate change on the country s natural resources could range between 1.0% and 4.8% of GDP (Reid et al., 2008). Ghana s agricultural and economic sector with cocoa being the single most important export product is particularly vulnerable, since cocoa is prone to the effects of a changing climate (Black et al., 2011c), which has been central to the country s debates on development and poverty alleviation strategies (WTO Trade Policy Review, 2008). The potential for adaptation to reduce the risks associated with sea level rise is substantial for cumulative land loss and for numbers of people flooded or forced to migrate, with adaptation costs lower than the economic and social damages expected if nothing is done (Kebede et al., 2010).See Figure 22-6. [INSERT FIGURE 22-6 HERE Figure 22-6: Total additional costs of adaptation per year from 2000 to 2100 for Tanzania (including beach nourishment and sea and river dikes). The values do not consider the existing adaptation deficit (values in $US 2005, without discounting Source: Kebede et al., 2010.] The Dynamic Interactive Vulnerability Assessment (DIVA) model was used to assess the monetary and non- monetary impacts of sea level rise on the entire coast (3,461 km) of Tanzania. Under the B1 low-range sea level rise scenario it was estimated that by 2030, a total area of 3,579 to 7,624 km2 would be lost, mainly through inundation; with around 234,000 to 1.6 million people per year who could potentially experience flooding. Without adaptation, residual damages have been estimated at between US$ 26 and 55 million per year (Kebede et al., 2010). Table 22-7 shows the economic impacts of land inundated in Cape Town based on different sea level rise scenarios. [INSERT TABLE 22-7 HERE Table 22-7: Land inundated and economic impacts in Cape Town based on a risk assessment (Cartwright, 2008).] In line with increasing international impetus for adaptation (Persson et al., 2009) the Parties to the UNFCCC agreed on providing adequate, predictable and sustainable financial resources for adaptation in developing countries, and, within this context, paid special attention to Africa which is particularly vulnerable to the adverse effects of climate change (UNFCCC, 2009; UNFCCC, 2011; Berenter, 2012). Doubts remain about how private sector financing can be effectively mobilized and channeled toward adaptation in developing countries (Atteridge, 2011; Naidoo et al., 2012). The 2012 Landscape of Climate Finance Report (Buchner et al., 2012) stated that mitigation activities attracted US$ 350 billion, mostly related to renewable energy and energy efficiency, while adaptation activities attracted US$ 14 billion. Approximately 30% of the global distributed adaptation finance went to Africa (Nakhooda et al., 2011) and seems to prioritize the continent (Naidoo et al., 2012). However, it is being questioned, whether the adaptation funding that is currently delivered does fulfill demonstrated needs (Flam and Skjaerseth, 2009; Denton, 2010 for sub-Saharan Africa Nakhooda et al., 2011). Effective adaptation requires more than sufficient levels of funding. It requires developing country readiness, which includes abilities to plan and access finances; the capacity to deliver adaptation projects and programs, and to monitor, report, and evaluate their effectiveness (Vandeweerd et al., 2012); and also a regulatory framework, which guarantees e.g. property rights (IPCC AR5 WGIII Draft 2 Chapter 16 p.:27; line 30-33). Particularly serious challenges are associated with directing finance to the sectors and people most vulnerable to climate change Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (Denton, 2010; Nakhooda et al., 2011; Pauw et al., 2012). The risk of fund mismanagement with regard to climate finance and adaptation funds needs to be borne in mind. Suggestions to address adequately the level of complexity, uncertainty, and novelty that surrounds many climate finance issues inter alia include longer-term and integrated programs rather than isolated projects; building capacity and institutions in African countries (Nakhooda et al., 2011; Pauw et al., 2012); identifying priorities, processes, and knowledge needs at the local level (Haites, 2011; Pauw, 2013); and, accordingly, developing grassroots projects (Fankhauser and Burton, 2011). 22.7. Research Gaps Research has a key role to play in providing information for informed decisionmaking at local to national levels (Fankhauser, 1997; Ziervogel et al., 2008; Arendse and Crane, 2010). While there is significant activity in African research institutions, much African research capacity is spent on foreign-led research that may necessarily prioritize addressing national knowledge gaps about climate change (Madzwamuse, 2010), and African research may lack merited policy uptake or global recognition as it is often not published in peer-reviewed literature (Denton et al., 2011). The following overarching data and research gaps have been identified also see Table 22.8: Data management and monitoring of climate and hydroclimate parameters and development of climate change scenarios as well as monitoring systems to address climate change impacts in the different sectors (for example the impacts of pests and diseases on crops and livestock) and systems; Research and improved methodologies to assess and quantify the impact of climate change on different sectors and systems. Socio-economic consequences of the loss of ecosystems and also of economic activities as well as of certain choices in terms of mitigation (biofuels and their links with food and livelihood security for example) and adaptation to climate change; The links influence of climate change in emerging issues such as migration and urban food security; Developing tools allowing decision makers to make their decisions based on the complexity of the world under climate change, taking into consideration gender, age and all regarding the contribution of local communities. [INSERT TABLE 22-8 HERE Table 22-8: Research gaps in different sectors.] Frequently Asked Questions FAQ 22.1: How could climate change impact food security in Africa? [to be inserted in Section 22.3.4.5] Food security is comprised of availability (is enough food produced), access (can people get it, and afford it), utilization (how local conditions bear on peoples nutritional uptake from food), and stability (is the supply and access ensured). Strong consensus exists that climate change will have a significantly negative impact on all these aspects of food security in Africa. Food availability could be threatened through direct climate impacts on crops and livestock from increased flooding, drought, shifts in the timing and amount of rainfall, and high temperatures, or indirectly through increased soil erosion from more frequent heavy storms or through increased pest and disease pressure on crops and livestock caused by warmer temperatures and other changes in climatic conditions. Food access could be threatened by climate change impacts on productivity in important cereal-producing regions of the world which, along with other factors, could raise food prices and erode the ability of the poor in Africa to afford purchased food. Access is also threatened by extreme events that impair food transport and other food system infrastructure. Climate change could impact food utilization through increased disease burden that reduces the ability of the human body to absorb nutrients from food. Warmer and more humid conditions caused by climate change could impact food availability and utilization through increased risk of spoilage of fresh food and pest and pathogen damage to stored foods (cereals, pulses, tubers) that reduces both food availability and quality. Stability could be affected by changes in availability and access that are linked to climatic and other factors. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 FAQ 22.2: What role does climate change play with regard to violent conflict in Africa? [to be inserted in Section 22.6.1.1] Wide consensus exists that violent conflicts are based on a variety of interconnected causes, of which the environment is considered to be one, but rarely the most decisive factor. Whether the changing climate increases the risk of civil war in Africa remains disputed and little robust research is available to resolve this question. Climate change impacts that intensify competition for increasingly scarce resources like freshwater and arable land, especially in the context of population growth, are areas of concern. The degradation of natural resources as a result of both overexploitation and climate change will contribute to increased conflicts over the distribution of these resources. In addition to these stressors, however, the outbreak of armed conflict depends on many country-specific sociopolitical, economic and cultural factors. Cross-Chapter Boxes Box CC-GC. Gender and Climate Change [Jon Barnett (Australia), Marta G. Rivera Ferre (Spain), Petra Tschakert (U.S.A.), Katharine Vincent (South Africa), Alistair Woodward (New Zealand)] Gender, along with socio-demographic factors of age, wealth and class, is critical to the ways in which climate change is experienced. There are significant gender dimensions to impacts, adaptation and vulnerability. This issue was raised in WGII AR4 and SREX reports (Adger et al., 2007; IPCC, 2012), but for the AR5 there are significant new findings, based on multiple lines of evidence on how climate change is differentiated by gender, and how climate change contributes to perpetuating existing gender inequalities. This new research has been undertaken in every region of the world (e.g. Brouwer et al., 2007; Nightingale, 2009; Buechler, 2009; Nelson and Stathers, 2009; Dankelman, 2010; MacGregor, 2010; Alston, 2011; Arora-Jonsson, 2011; Resureccion, 2011; Omolo, 2011). Gender dimensions of vulnerability derive from differential access to the social and environmental resources required for adaptation. In many rural economies and resource-based livelihood systems, it is well established that women have poorer access than men to financial resources, land, education, health and other basic rights. Further drivers of gender inequality stem from social exclusion from decision-making processes and labour markets, making women in particular less able to cope with and adapt to climate change impacts (Rijkers and Costa, 2012; Djoudi and Brockhaus, 2011; Paavola, 2008). These gender inequalities manifest themselves in gendered livelihood impacts and feminisation of responsibilities: whilst both men and women experience increases in productive roles, only women experience increased reproductive roles (Resureccion, 2011; 9.3.5.1.5, Box 13-1). A study in Australia, for example, showed how more regular occurrence of drought has put women under increasing pressure to earn off-farm income, and contribute to more on-farm labor (Alston, 2011). Studies in Tanzania and Malawi demonstrate how women experience food and nutrition insecurity since food is preferentially distributed among other family members (Nelson and Stathers, 2009; Kakota et al., 2011). AR4 assessed a body of literature that focused on women s relatively higher vulnerability to weather-related disasters in terms of number of deaths (Adger et al., 2007). Additional literature published since that time adds nuances by showing how socially-constructed gender differences affect exposure to extreme events, leading to differential patterns of mortality for both men and women (high confidence) [11.3.3, Table 12-3]. Statistical evidence of patterns of male and female mortality from recorded extreme events in 141 countries between 1981- 2002 found that disasters kill women at an earlier age than men (Neumayer and Plümper, 2007) [Box 13-1]. Reasons for gendered differences in mortality include various socially- and culturally-determined gender roles. Studies in Bangladesh, for example, show that women do not learn to swim and so are vulnerable when exposed to flooding (Röhr, 2006) and that, in Nicaragua, the construction of gender roles means that middle-class women are expected to stay in the house, even during floods and in risk-prone areas (Bradshaw, 2010). While the differential vulnerability of women to extreme events has long been understood, there is now increasing evidence to show how gender roles for men can affect their vulnerability. In particular, men are often expected to be brave and heroic, and engage in risky life-saving behaviors that increase their likelihood of mortality [Box 13-1]. In Hai Lang district, Vietnam, for example, more men died than women due to their involvement in search and rescue and protection of fields during Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 flooding (Campbell et al., 2009). Women and girls are more likely to become victims of domestic violence after a disaster, particularly when they are living in emergency accommodation, which has been documented in the U.S. and Australia (Jenkins and Phillips, 2008; Anastario et al., 2009; Alston, 2011; Whittenbury, 2013; Box 13-1). Heat stress exhibits gendered differences, reflecting both physiological and social factors (11.3.3). The majority of studies in European countries show women to be more at risk, but their usually higher physiological vulnerability can be offset in some circumstances by relatively lower social vulnerability (if they are well connected in supportive social networks, for example). During the Paris heat wave, unmarried men were at greater risk than unmarried women, and in Chicago elderly men were at greatest risk, thought to reflect their lack of connectedness in social support networks which led to higher social vulnerability (Kovats and Hajat, 2008). A multi-city study showed geographical variations in the relationship between sex and mortality due to heat stress: in Mexico City, women had a higher risk of mortality than men, although the reverse was true in Santiago and Sao Paulo (Bell et al., 2008). Recognizing gender differences in vulnerability and adaptation can enable gender-sensitive responses that reduce the vulnerability of women and men (Alston, 2013). Evaluations of adaptation investments demonstrate that those approaches that are not sensitive to gender dimensions and other drivers of social inequalities risk reinforcing existing vulnerabilities (Figueiredo and Perkins, 2012; Arora-Jonsson, 2011; Vincent et al., 2010). Government- supported interventions to improve production through cash-cropping and non-farm enterprises in rural economies, for example, typically advantage men over women since cash generation is seen as a male activity in rural areas (Gladwin et al., 2001;13.3.1). In contrast, rainwater and conservation-based adaptation initiatives may require additional labor which women cannot necessarily afford to provide (Baiphethi et al., 2008). Encouraging gender- equitable access to education and strengthening of social capital are among the best means of improving adaptation of rural women farmers (Below et al., 2012; Goulden et al., 2009; Vincent et al., 2010) and could be used to complement existing initiatives mentioned above that benefit men. Rights-based approaches to development can inform adaptation efforts as they focus on addressing the ways in which institutional practices shape access to resources and control over decision-making processes, including through the social construction of gender and its intersection with other factors that shape inequalities and vulnerabilities (Tschakert, 2013; Bee et al., 2013; Tschakert and Machado, 2012; see also 22.4.3 and Table 22-5). Box CC-GC References Adger, W.N., S. Agrawala, M.M.Q. Mirza, C. Conde, K. O Brien, J. Pulhin, R. Pulwarty, B. Smit, and K. Takahashi, 2007: Chapter 17: Assessment of adaptation practices, options, constraints and capacity. In: Climate Change 2007: Synthesis Report. Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [IPCC (ed.)]. IPCC, Geneva, Switzerland, pp. 719-743. Alston, M., 2011: Gender and climate change in Australia. Journal of Sociology, 47(1), 53-70. Alston, M., 2013: Women and adaptation. Wiley Interdisciplinary Reviews: Climate Change, (4)5, 351-358. Anastario, M., N. Shebab, and L. Lawry, 2009: Increased gender-based violence among women internally displaced in Mississippi 2 years post- Hurricane Katrina. Disaster Medicine and Public Health Preparedness, 3(1), 18-26. Arora-Jonsson, S., 2011: Virtue and vulnerability: Discourses on women, gender and climate change. Global Environmental Change, 21, 744- 751. Baiphethi, M.N., M. Viljoen, and G. Kundhlande, 2008: Rural women and rainwater harvesting and conservation practices: Anecdotal evidence from the Free State and Eastern Cape. Agenda, 22(78), 163-171. Bee, B., M. Biermann, and P. Tschakert, 2013: Gender, development, and rights-based approaches: Lessons for climate change adaptation and adaptive social protection. In: Research, Action and Policy: Addressing the Gendered Impacts of Climate Change. [Alston, M. and K. Whittenbury(eds.)]. Springer, Netherlands, pp. 95-108. Bell M.L., M.S. O'Neill, N. Ranjit, V.H. Borja-Aburto, L.A. Cifuentes and N.C. Gouveia, 2008: Vulnerability to heat-related mortality in Latin America: a case-crossover study in Sao Paulo, Brazil, Santiago, Chile and Mexico City, Mexico. International Journal of Epidemiology 37(4), 796 804. Below, T.B., K.D. Mutabazi, D. Kirschke, C. Franke, S. Sieber, R. Siebert, and K. Tscherning, 2012: Can farmers adaptation to climate change be explained by socio-economic household-level variables? Global Environmental Change, 22(1), 223-235. Bradshaw, S., 2010: Women, poverty, and disasters: Exploring the links through hurricane Mitch in Nicaragua. In: The international handbook of gender and poverty: concepts, research, policy. [Chant, S. (ed.)]. Edward Elgar Pub, Cheltenham, UK, pp. 627. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Brouwer, R., S. Akter, L. Brander, and E. Haque, 2007: Socioeconomic vulnerability and adaptation to environmental risk: A case study of climate change and flooding in Bangladesh. Risk Analysis, 27(2), 313-326. Campbell, B., S. Mitchell, and M. Blackett, 2009: Responding to Climate Change in Vietnam. Opportunities for Improving Gender Equality. Oxfam; UNDP, Hanoi, Vietnam, pp. 1-63. Dankelman, I., 2010: Introduction: Exploring gender, environment, and climate change. In: Gender and climate change: An introduction. [Dankelman, I. (ed.)]. Earthscan, London, UK, Sterling, VA, USA, pp. 1-20. Djoudi, H. and M. Brockhaus, 2011: Is adaptation to climate change gender neutral? Lessons from communities dependent on livestock and forests in northern Mali. International Forestry Review, 13(2), 123-135. Figueiredo, P. and P.E. Perkins, 2012: Women and water management in times of climate change: participatory and inclusive processes. Journal of Cleaner Production, (online). Gladwin, C.H., A.M. Thomson, J.S. Peterson, and A.S. Anderson, 2001: Addressing food security in Africa via multiple livelihood strategies of women farmers. Food Policy, 26(2), 177-207. Goulden, M., L.O. Naess, K. Vincent, and W.N. Adger, 2009: Diversification, networks and traditional resource management as adaptations to climate extremes in rural Africa: opportunities and barriers. In: Adapting to Climate Change: Thresholds, Values and Governance. [Adger, W.N., I. Lorenzoni, and K. O Brien(eds.)]. Cambridge University Press, Cambridge, pp. 448-464. IPCC (ed.), 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change. Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley, Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 582. Jenkins, P. and B. Phillips, 2008: Battered Women, Catastrophe, and the Context of Safety after Hurricane Katrina. NWSA, 20(3), 49-68. Kakota, T., D. Nyariki, D. Mkwambisi, and W. Kogi-Makau, 2011: Gender vulnerability to climate variability and household food insecurity. Climate and Development, 3(4), 298-309. Kovats R, Hajat S., 2008: Heat stress and public health: a critical review. Public Health, 29, 41-55. MacGregor, S., 2010: Gender and climate change : from impacts to discourses. Journal of the Indian Ocean Region, 6(2), 223-238. Nelson, V. and T. Stathers, 2009: Resilience, power, culture, and climate: a case study from semi-arid Tanzania, and new research directions. Gender & Development, 17(1), 81-94. Neumayer, E. and T. Plümper, 2007: The gendered nature of natural disasters: The impact of catastrophic events on the gender gap in life expectancy, 1981 2002. Annals of the Association of American Geographers, 97(3), 551-566. Nightingale, A., 2009: Warming up the climate change debate: A challenge to policy based on adaptation. Journal of Forest and Livelihood, 8(1), 84-89. Omolo, N., 2011: Gender and climate change-induced conflict in pastoral communities: Case study of Turkana in northwestern Kenya. African Journal on Conflict Resolution, 10(2), 81-102. Paavola, J., 2008: Livelihoods, vulnerability and adaptation to climate change in Morogoro, Tanzania. Environmental Science & Policy, 11(7), 642-654. Resurreccion, B.P., 2011: The Gender and Climate Debate: More of the Same or New Pathways of Thinking and Doing?. In: Asia Security Initiative Policy Series. RSIS Centre for Non-Traditional Security (NTS) Studies, Singapore, pp. 1-22. Rijkers, B. and Costa, R., 2012: Gender and Rural Non-Farm Entrepreneurship, Policy research working papers, 6066, World Bank, pp. 68 Röhr, U., 2006: Gender and climate change. Tiempo, 59, 3-7. Tschakert, P., 2013: From impacts to embodied experiences: tracing political ecology in climate change research, Geografisk Tidsskrift-Danish Journal of Geography, 112(2), 144-158. Tschakert, P. and M. Machado, 2012: Gender Justice and Rights in Climate Change Adaptation: Opportunities and Pitfalls., Ethics and Social Welfare, doi: 10.1080/17496535.2012.704929. Vincent, K., T. Cull, and E. Archer, 2010: Gendered vulnerability to climate change in Limpopo province, South Africa. In: Gender and Climate Change: An Introduction. [Dankelman, I. (ed.)]. Earthscan, London, pp. 160-167. Whittenbury, K., 2013: Climate Change, Women's Health, Wellbeing and Experiences of Gender-Based Violence in Australia. In: Research, Action and Policy: Addressing the Gendered Impacts of Climate Change. [Alston, M. and K. Whittenbury(eds.)]. Springer, Australia, pp. 207-222. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Box CC-UP. Uncertain Trends in Major Upwelling Ecosystems [Salvador E. Lluch-Cota (Mexico), Ove Hoegh-Guldberg (Australia), David Karl (USA), Hans O. Pörtner (Germany), Svein Sundby (Norway), Jean-Pierre Gatusso (France)] Upwelling is the vertical transport of cold, dense, nutrient-rich, relatively low-pH and often oxygen-poor waters to the euphotic zone where light is abundant. These waters trigger high levels of primary production and a high biomass of benthic and pelagic organisms. The driving forces of upwelling include wind stress and the interaction of ocean currents with bottom topography. Upwelling intensity also depends on water column stratification. The major upwelling systems of the Planet, the Equatorial Upwelling System (EUS, 30.5.2, Figure 30.1A) and the Eastern Boundary Upwelling Ecosystems (EBUE, 30.5.5, Figure 30.1A), represent only 10% of the ocean surface but contribute nearly 25 % to global fish production (Figure 30.1B, Table S30.1). Marine ecosystems associated with upwelling systems can be influenced by a range of bottom-up trophic mechanisms, with upwelling, transport, and chlorophyll concentrations showing strong seasonal and interannual couplings and variability. These, in turn, influence trophic transfer up the food chain, affecting zooplankton, foraging fish, seabirds and marine mammals. There is considerable speculation as to how upwelling systems might change in a warming and acidifying ocean. Globally, the heat gain of the surface ocean has increased stratification by 4% (WGI 3.2, 3.4.4, 3.8), which means that more wind energy is needed to bring deep waters to the surface. It is as yet unclear to what extent wind stress can offset the increased stratification, due to the uncertainty in wind speed trends (WGI, 3.4.4). In the tropics, observations of reductions in trade winds over several decades contrast more recent evidence indicating their strengthening since the early 1990s (WGI, 9.4.1.3.4). Observations and modelling efforts in fact show diverging trends in coastal upwelling at the eastern boundaries of the Pacific and the Atlantic. Bakun (1990) proposed that the the difference in heat gaining rates between land and ocean causes an increase in the pressure gradient, which results in increased alongshore winds and leads to intensified offshore transport of surface water through Ekman pumping, and the upwelling of nutrient rich, cold waters (Figure CC-UP). Some regional records support this hypothesis, others do not. There is considerable variability in warming and cooling trends over the past decades both within and among systems making it difficult to predict changes in the intensity of all Eastern Boundary Upwelling Ecosystems (30.5.5). Understanding whether upwelling and climate change will impact resident biota in an additive, synergistic or antagonistic manner is important for projections of how ecological goods and services provided for human society will change. Even though upwellings may prove more resilient to climate change than other ocean ecosystems because of their ability to function under extremely variable conditions (Capone and Hutchins, 2013), consequences of their shifts are highly relevant since these are the most biologically active systems in the ocean. Increased upwelling would enhance fisheries yields. However, the export of organic material from surface to deeper layers of the ocean may increase and stimulate its decomposition by microbial activity, thereby enhancing oxygen depletion and CO2 enrichment in deeper water layers. Once this water returns to the surface through upwelling benthic and pelagic coastal communities will be exposed to acidified and deoxygenated water which may combine with anthropogenic impact to negatively affect marine biota and ecosystem structure of the upper ocean (high confidence, 6.3.2, 6.3.3; 30.3.2.2, 30.3.2.3). Extreme hypoxia may result in abnormal mortalities of fishes and invertebrates (Keller et al., 2010), reduce the fisheries catch potential and impact aquaculture in coastal areas (5.4.3.3, 6.3.7, 30.5.1.1.2, 30.5.5.1.3, Barton et al., 2012). Shifts in upwelling also coincide with an apparent increase in the frequency of submarine eruptions of methane and hydrogen sulphide gas, caused by enhanced formation and sinking of phytoplankton biomass to the hypoxic or anoxic sea floor . This combination of factors has been implicated in the extensive mortality of coastal fishes and invertebrates (Bakun and Weeks, 2004), resulting in significant reductions in fishing productivity, such as Cape hake (Merluccius capensis), Namibia s most valuable fishery (Hamukuaya et al., 1998). Reduced upwelling would also reduce the productivity of important pelagic fisheries, such as for sardines, anchovies and mackerel, with major consequences for the economies of several countries (6.4.1, Chp 7, Figure 30.1A, B, Table S30.1). However, under projected scenarios of reduced upward supply of nutrients due to stratification of the open ocean , upwelling of both nutrients and trace elements may become increasingly important to maintaining upper Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 ocean nutrient and trace metal inventories. It has been suggested that upwelling areas may also increase nutrient content and productivity under enhanced stratification, and that upwelled and partially denitrified waters containing excess phosphate may select for N2-fixing microorganisms (Deutsch et al., 2007; Deutsch and Weber, 2012), but field observations of N2 fixation in these regions have not supported these predictions (Fernandez et al., 2011; Franz et al., 2012). The role of this process in global primary production thus needs to be validated (low confidence). The central question therefore is whether or not upwelling will intensify, and if so, whether the effects of intensified upwelling on O2 and CO2 inventories will outweigh its benefits for primary production and associated fisheries and aquaculture (low confidence). In any case increasing atmospheric CO2 concentrations will equilibrate with upwelling waters that may cause them to become more corrosive, depending upon pCO2 of the upwelled water, and potentially increasingly impact the biota of Eastern Boundary Upwelling Ecosystems. [INSERT FIGURE UP-1 HERE Figure UP-1: Upper panel: Schematic hypothetic mechanism of increasing coastal wind-driven upwelling at eastern boundary systems, where differential warming rates between land and ocean results in increased land-ocean pressure gradients (1) that produce stronger alongshore winds (2) and offshore movement of surface water through Ekman transport (3), and increased upwelling of deep cold nutrient rich waters to replace it (4). Lower panel: potential consequences of climate change in upwelling systems. Increasing stratification and uncertainty in wind stress trends result in uncertain trends in upwelling. Increasing upwelling may result in higher input of nutrients to the euphotic zone, and increased primary production, which in turn may enhance pelagic fisheries, but also decreased coastal fisheries due to an augmented exposure of coastal fauna to hypoxic, low pH waters. Decreased upwelling may result in lower primary production in these systems with direct impacts on pelagic fisheries productivity.] Box CC-UP References Bakun, A., 1990: Global climate change and intensification of coastal ocean upwelling, Science, 247(4939), 198-201. Bakun, A. and S.J. Weeks, 2004: Greenhouse gas buildup, sardines, submarine eruptions and the possibility of abrupt degradation of intense marine upwelling ecosystems. Ecology Letters, 7(11), 1015-1023. Barton, A., B. Hales, G.G. Waldbusser, C. Langdon, R.A. Feely, 2012: The Pacific oyster, Crassostrea gigas, shows negative correlation to naturally elevated carbon dioxide levels: Implications for near-term ocean acidification effects, Limnology and Oceanography, 57(3): 698- 710. Capone, D.G. and D.A. Hutchins, 2013: Microbial biogeochemistry of coastal upwelling regimes in a changing ocean. Nature geoscience, 711- 717. Deutsch, C. and T. Weber, 2012: Nutrient ratios as a tracer and driver of ocean biogeochemistry. Annual Review of Marine Science, 4, 113-114. Deutsch, C., J.L. Sarmiento, D.M. Sigman, N. Gruber and J.P. Dunne, 2007: Spatial coupling of nitrogen inputs and losses in the ocean. Nature, 445(7124), 163-167. Fernandez, C., L. Farías and O. Ulloa, 2011: Nitrogen fixation in denitrified marine waters. PLoS ONE, 6(6), e20539. Franz, J., G. Krahmann, G. Lavik, P. Grasse, T. Dittmar and U. Riebesell, 2012: Dynamics and stoichiometry of nutrients and phytoplankton in waters influenced by the oxygen minimum zone in the eastern tropical Pacific. Deep Sea Research Part I: Oceanographic Research Papers, 62, 20-31. Hamukuaya, H., M.J. O'Toole and P.M.J. Woodhead, 1998: Observations of severe hypoxia and offshore displacement of Cape hake over the Namibian shelf in 1994. South African Journal of Marine Science, 19(1), 57-59. Keller, AA, Simon V, Chan F, Wakefield WW, Clarke ME, et al., 2010: Demersal fish and invertebrate biomass in relation to an offshore hypoxic zone along the US West Coast. Fisheries Oceanography 19:76 87. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Box CC-WE: The Water-Energy-Food/Feed/Fiber Nexus as Linked to Climate Change [Douglas J. Arent (USA), Petra Döll (Germany), Kenneth M. Strzepek (UNU / USA), Blanca Elena Jimenez Cisneros (Mexico), Andy Reisinger (New Zealand), FerencToth (IAEA / Hungary), Taikan Oki (Japan)] Water, energy, and food/feed/fiber are linked through numerous interactive pathways and subject to a changing climate, as depicted in Figure WE-1. The depth and intensity of those linkages vary enormously between countries, regions and production systems. Energy technologies (e.g. biofuels, hydropower, thermal power plants), transportation fuels and modes, and food products (from irrigated crops, in particular animal protein produced by feeding irrigated crops and forages) may require significant amounts of water (Sections 3.7.2, 7.3.2, 10.2, 10.3.4, 22.3.3, 25.7.2; Allan, 2003; King and Weber 2008; McMahon and Price, 2011; Macknick et al., 2012a). In irrigated agriculture, climate, irrigating procedure, crop choice and yields determine water requirements per unit of produced crop. In areas where water (and wastewater) must be pumped and/or treated, energy must be provided (Asano et al., 2006; Khan and Hanjra, 2009; USEPA, 2010; Gerten et al., 2011). While food production, refrigeration, transport and processing require large amounts of energy (Pelletier et al., 2011), a major link between food and energy as related to climate change is the competition of bioenergy and food production for land and water (Section 7.3.2, Box 25-10; Diffenbaugh et al., 2012; Skaggs et al., 2012) (robust evidence, high agreement). Food and crop wastes, and wastewater, may be used as sources of energy, saving not only the consumption of conventional non-renewable fuels used in their traditional processes, but also the consumption of the water and energy employed for processing or treatment and disposal (Schievano et al., 2009; Sung et al., 2010; Olson, 2012). Examples of this can be found in several countries across all income ranges. For example, sugar cane by-products are increasingly used to produce electricity or for cogeneration (McKendry, 2002; Kim and Dale 2004) for economic benefits, and increasingly as an option for greenhouse gas mitigation. [INSERT FIGURE WE-1 HERE Figure WE-1: The water-energy-food nexus as related to climate change. The interlinkages of supply/demand, quality and quantity of water, energy and food/feed/fiber with changing climatic conditions have implications for both adaptation and mitigation strategies.] Most energy production methods require significant amounts of water, either directly (e.g., crop-based energy sources and hydropower) or indirectly (e.g., cooling for thermal energy sources or other operations) (Sections 10.2.2 10.3.4, 25.7.4; van Vliet et al., 2012; Davies et al., 2013) (robust evidence, high agreement). Water for biofuels, for example, under the IEA Alternative Policy Scenario, which has biofuels production increasing to 71 EJ in 2030, has been reported by Gerbens-Leenes et al. (2012) to drive global consumptive irrigation water use from 0.5% of global renewable water resources in 2005 to 5.5% in 2030, resulting in increased pressure on freshwater resources, with potential negative impacts on freshwater ecosystems. Water is also required for mining (Section 25.7.3), processing, and residue disposal of fossil and nuclear fuels or their byproducts. Water for energy currently ranges from a few percent in most developing countries to more than 50% of freshwater withdrawals in some developed countries, depending on the country (Kenny et al., 2009; WEC, 2010). Future water requirements will depend on electricity demand growth, the portfolio of generation technologies and water management options employed (WEC, 2010; Sattler et al., 2012) (medium evidence, high agreement). Future water availability for energy production will change due to climate change (Sections 3.4, 3.5.1, 3.5.2.2) (robust evidence, high agreement). Water may require significant amounts of energy for lifting, transport and distribution and for its treatment either to use it or to depollute it. Wastewater and even excess rainfall in cities requires energy to be treated or disposed. Some non-conventional water sources (wastewater or seawater) are often highly energy intensive. Energy intensities per m3 of water vary by about a factor of 10 between different sources, e.g. locally produced potable water from ground/surface water sources vs. desalinated seawater (Box 25-2, Tables 25-6 and 25-7; Macknick et al., 2012b; Plappally and Lienhard, 2012). Groundwater (35% of total global water withdrawals, with irrigated food production being the largest user; Döll et al., 2012) is generally more energy intensive than surface water. In India, for example, 19% of total electricity use in 2012 was for agricultural purposes (Central Statistics Office, 2013), with a large share for groundwater pumping. Pumping from greater depth increases energy demand significantly electricity use (kWhr/m3 of water) increases by a factor of 3 when going from 35 to 120 m depth (Plappally and Lienhard, 2012). The reuse of appropriate wastewater for irrigation (reclaiming both water and energy-intense nutrients) may increase agricultural yields, save energy, and prevent soil erosion (Smit and Nasr, 1992; Jimenez, 1996; Wichelns et al., Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2007; Raschid-Sally and Jayakody, 2008) (medium confidence). More energy efficient treatment methods enable poor quality ( black ) wastewater to be treated to quality levels suitable for discharge into water courses, avoiding additional fresh water and associated energy demands (Keraita et al, 2008). If properly treated to retain nutrients, such treated water may increase soil productivity, contributing to increased crop yields/food security in regions unable to afford high power bills or expensive fertilizer (Oron, 1996; Lazarova and Bahri, 2005; Redwood and Huibers, 2008; Jimenez, 2009) (high confidence). Linkages among water, energy, food/feed/fiber and climate are also strongly related to land use and management (Section 4.4.4, Box 25-10) (robust evidence, high agreement). Land degradation often reduces efficiency of water and energy use (e.g. resulting in higher fertilizer demand and surface runoff), and compromises food security (Sections 3.7.2, 4.4.4). On the other hand, afforestation activities to sequester carbon have important co-benefits of reducing soil erosion and providing additional (even if only temporary) habitat (see Box 25-10) but may reduce renewable water resources. Water abstraction for energy, food or biofuel production or carbon sequestration can also compete with minimal environmental flows needed to maintain riverine habitats and wetlands, implying a potential conflict between economic and other valuations and uses of water (Sections 25.4.3 and 25.6.2, Box 25-10) (medium evidence, high agreement). Only a few reports have begun to evaluate the multiple interactions among energy, food, land, and water and climate (McCornick et al., 2008; Bazilian et al., 2011; Bierbaum and Matson, 2013), addressing the issues from a security standpoint and describing early integrated modeling approaches. The interaction among each of these factors is influenced by the changing climate, which in turn impacts energy and water demand, bioproductivity and other factors (see Figure WE-1 and Wise et al., 2009), and has implications for security of supplies of energy, food and water, adaptation and mitigation pathways, air pollution reduction as well as the implications for health and economic impacts as described throughout this report. The interconnectivity of food/fiber, water, land use, energy and climate change, including the perhaps not yet well understood cross-sector impacts, are increasingly important in assessing the implications for adaptation/mitigation policy decisions. Fuel-food-land use-water-GHG mitigation strategy interactions, particularly related to bioresources for food/feed, power, or fuel, suggest that combined assessment of water, land type and use requirements, energy requirements and potential uses and GHG impacts often epitomize the interlinkages. For example, mitigation scenarios described in the IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation (IPCC, 2011) indicate up to 300EJ of biomass primary energy by 2050 under increasingly stringent mitigation scenarios. Such high levels of biomass production, in the absence of technology and process/management/operations change, would have significant implications for land use, water and energy, as well as food production and pricing. Consideration of the interlinkages of energy, food/feed/fiber, water, land use and climate change is increasingly recognized as critical to effective climate resilient pathway decision making (medium evidence, high agreement), although tools to support local- and regional-scale assessments and decision-support remain very limited. Box CC-WE References Asano T., Burton F., Leverenz H, Tsuchihashi R., Tchobanoglous F., 2006: Water Reuse: Issues, Technologies, and Applications, Metcalf & Eddy, Inc. An AECOM Company, New York, 1570 pp Allan T., 2003: Virtual Water - the Water, Food, and Trade Nexus Useful Concept or Misleading Metaphor? IWRA, Water International; 28(1), 4-10. Bazilian, M. Rogner, H., Howells, M., Hermann, S., Arent, D., Gielen, D., Steduto, P., Mueller, A., Komor, P., Tol, R.S.J., Yumkella, K., 2011: Considering the energy, water and food nexus: Towards an integrated modelling approach. Energy Policy, 39(12), 7896-7906. Bierbaum, R., and P. Matson, 2013: Energy in the Context of Sustainability, Daedalus, The Alternative Energy Future. 2, 90-97. Davies, E., Page, K. and Edmonds, J. A., 2013: An Integrated Assessment of Global and Regional Water Demands for Electricity Generation to 2095." Advances in Water Resources 52, 296 313.10.1016/j.advwatres.2012.11.020. Diffenbaugh, N.,Hertel, T., M. Scherer & M. Verma, 2012: Response of corn markets to climate volatility under alternative energy futures. Nature Climate Change 2, 514 518. Döll, P., Hoffmann-Dobrev, H., Portmann, F.T., Siebert, S., Eicker, A., Rodell, M., Strassberg, G., Scanlon, B., 2012: Impact of water withdrawals from groundwater and surface water on continental water storage variations. J. Geodyn. 59-60, 143-156, doi:10.1016/j.jog.2011.05.001. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Gerbens-Leenes, P.W., A.R. van Lienden, A.Y. Hoekstra, Th.H. van der Meer, 2012: Biofuel scenarios in a water perspective: The global blue and green water footprint of road transport in 2030. Global Environmental Change, 22(3), 764-775. Gerber, N., M. van Eckert and T. Breuer, 2008: The Impacts of Biofuel Production on Food Prices: a review, ZEF Discussion Papers On Development Policy, No. 127, Center for Development Research, Bonn, pp.19. Gerten D., Heinke H., Hoff H., Biemans H., Fader M., Waha K., 2011: Global water availability and requirements for future food production, Journal of Hydrometeorology, doi: 10.1175/2011JHM1328.1. IPCC, 2011: Summary for Policymakers. In: IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation; [O. Edenhofer, R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner, T. Zwickel, P. Eickemeier, G. Hansen, S. Schlmer, C. von Stechow (eds)], Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Jiménez, B., 1996: Wastewater Reuse to Increase Soil Productivity. Water Science and Technology, 32(12), 173-180. Jiménez, B., 2009: Safe Sanitation in Low Economic Development Areas, Treatise MS 82. in Treatise on Water, Science. Vol. 4, [Wilderer P. ed.], Oxford, Academic Press. pp. 147 201. Kenny, J.F., Barber, N.L., Hutson, S.S., Linsey, K.S., Lovelace, J.K., and Maupin, M.A.,2009: Estimated use of water in the United States in 2005. U.S. Geological Survey Circular 1344, 52p. Keraita B., Jiménez B. and Drechsel P., 2008: Extent and Implications of Agricultural Reuse of Untreated, partly Treated and Diluted Wastewater in Developing Countries. CAB Reviews: Perspectives in Agriculture, Veterinary Science, Nutrition and Natural Resources, 3(58), 15-27. Khan, S., Hanjra, M. A. 2009: Footprints of water and energy inputs in food production - Global perspectives. Food Policy, 34, 130-140. King, C. and Webber, M. E., Water intensity of transportation, Environmental Science and Technology, 2008: 42 (21), 7866-7872. Kim, S. and Dale B., 2004: Global potential bioethanol production from wasted crops and crop residues. Biomass and Bioenergy 26(4), 361 375. Lazarova, V., A. Bahri, 2005: Water reuse for irrigation: agriculture, landscapes, and turf grass, CRC Press, Boca Raton, FL, 408 pp. McMahon, J.E. and Sarah K. Price, 2011: Water and Energy Interactions; Annu. Rev. Environ. Resour. 36, 163 91. McKendry P., 2002: Energy production from biomass (part 1): overview of biomass. Bioresource Technology , 83, 37 46. Macknick, J.; Newmark, R.; Heath, G.; Hallett, K. C.; Meldrum, J.; Nettles-Anderson, S., 2012a:. Operational Water Consumption and Withdrawal Factors for Electricity Generating Technologies: A Review of Existing Literature. Environmental Research Letters. 7(4). Macknick, J.; S. Sattler, K. Averyt, S. Clemmer, J. Rogers, 2012b: Water Implications of Generating Electricity: Water Use Across the United States Based on Different Electricity Pathways through 2050. Environmental Research Letters. 7(4). McCornick P.G., S.B. Awulachew. and Abebe M., 2008: Water-food-energy-environment synergies and tradeoffs: major issues and case studies. Water Policy, 10, 23-36. Olson G.,2012: Water and energy Nexus: Threats and Opportunities, IWAP, London, 300 pp. Oron G.. 1996: Soil as a complementary treatment component for simultaneous wastewater disposal and reuse. Water Science and Technology, 34(11), 243-252. Plappally, A.K., and J.H. Lienhard V., 2012: Energy requirements for water production, treatment, end use, reclamation, and disposal. Renewable and Sustainable Energy Reviews, 16(7), 4818-4848. Pelletier, N., E. Audsley, S. Brodt, T. Garnett, P. Henriksson, A. Kendall, K.J. Kramer, D. Murphy, T. Nemeck, and M. Troell, 2011: Energy Intensity of Agriculture and Food Systems, Annual Review of Environment and Resources, 36 223-246. Raschid-Sally, L, P. Jayakody, 2008: Drivers and characteristics of wastewater agriculture in developing countries: Results from a Global Assessment, IWMI Research International Water Management Institute, Colombo, Sri Lanka, 127, p35. Redwood, M, and Huibers F, 2008: Wastewater irrigation in Urban Agriculture. In: Water Reuse An International Survey of current practices, issues and need, [Jimenez B and Asano T. (eds.)] IWAP London. pp.628. Sattler, S., J. Macknick, D. Yates, F. Flores-Lopez, A. Lopez, J. Rogers, 2012: Linking Electricity and Water Models to Assess Electricity Choices at Water-Relevant Scales. Environmental Research Letters. 7(4). Schievano A., G. D Imporzano, F. Adani , 2009: Substituting energy crops with organic wastes and agro-industrial residues for biogas production. Journal of Environmental Management 90, 2537 2541. Smit J, J. Nasr, 1992: Urban agriculture for sustainable cities: using wastes and idle land and water bodies as resources. Environment and Urbanization. 4(2), 141-52. Skaggs, R., T.C. Janetos, K.A. Hibbard, J.S. Rice, 2012: Climate and Energy-Water-Land System Interactions; Technical Report to the U.S. Department of Energy in Support of the National Climate Assessment, PNNL report 21185. Sung T. Oh, Jung Rae Kim Giuliano C. Premier, Tae Ho Lee, Changwon Kim, William T. Sloan, 2010: Sustainable wastewater treatment: How might microbial fuel cells contribute. 6, s.l. : Biotechnology Advances. November---December 2010, 28. USEPA, 2010: Evaluation of Energy Conservation Measures for Wastewater Treatment Facilities. EPA 832-R-10-005 , Washington, D.C., 222 pp, http://water.epa.gov/scitech/wastetech/upload/Evaluation-of-Energy-Conservation-Measures-for-Wastewater-Treatment-Facilities.pdf Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 van Vliet, M.T.H., , J.R., Ludwig, F., Vögele, S., Lettenmaier, D. P., and Kabat, P. , Vulnerability of US and European electricity supply to climate change. Nature Climate Change, 2, 676 681(2012). Wichelns D., L. Raschid-Sally, P. Singh Minhas, P. Drechsel, A. Bahri and P. McCornick (leading authors), R. Abaidoo, F. Attia, S. El Guindy, L. Ensink, B. Jiménez, J. Kijne, S. Koo-Oshima, J.D. Oster, L. Oyebande, J. A. Sagardoy and W. Van der Hoek, 2007: Agricultural Use of Marginal-quality Water-opportunities and Challenges. In Water for Food, Water for Life-A Comprehensive Assessment of Water Management in Agriculture. [Molden D. Ed.], Earthscan Publications Ltd. London, UK, pp. 425-458. Wise, K. Calvin, A. Thomson, L. Clarke, B. Bond-Lamberty, R. Sands, S.J. Smith, A. Janetos, J. Edmonds, 2009: Implications of limiting CO2 concentrations for land use and energy. Science 324, 1183-1186. World Energy Council (WEC); Water for Energy; 2010. References AAP, 2012: AAP knowledge management needs survey report 2012. United Nations Development Programme (UNDP) Africa Adaptation Programme (AAP), pp. 110. Abdo, K.S., B.M. Fiseha, T.H.M. Rientjes, A.S.M. Gieske, and A.T. Haile, 2009: Assessment of climate change impacts on the hydrology of Gilgel Abay catchment in Lake Tana basin, Ethiopia. Hydrological Processes, 23(26), 3661-3669. Abouabdillah, A., O. Oueslati, A.M. De Girolamo, and A. Lo Porto, 2010: Modeling the impact of climate change in a mediterranean catchment (Merguellil, Tunisia). Fresenius Environmental Bulletin, 19(10 A), 2334-2347. Abou-Hadid, A.F., 2006: Assessment of Impacts, Adaptation, and Vulnerability to Climate Change in North Africa: Food Production and Water Resources. A Final Report Submitted to Assessments of Impacts and Adaptations to Climate Change (AIACC), Project No. AF 90. International START Secretariat, Washington, DC, pp. 148. Abramovitz, J., T. Banuri, P.O. Girot, B. Orlando, N. Schneider, E. Spanger-Siegfried, J. Switzer, and A. Hammill, 2002: Adapting to Climate Change: Natural Resource Management and Vulnerability Reduction. Background Paper to the Task Force on Climate Change, Adaptation and Vulnerable Communities. World Conservation Union (IUCN), Worldwatch Institute, International Institute for Sustainable Development (IISD) and Stockholm Environment Institute /Boston (SEI-B), pp. 37. ACCRA, 2012: From local to national: Supporting local government action in climate change adaptation, disaster risk reduction and development planning. ACCRA Briefs. Africa Climate Change Resilience Alliance (ACCRA), pp. 4. Achten, W.M.J., L. Verchot, Y.J. Franken, E. Mathijs, V.P. Singh, R. Aerts, and B. Muys, 2008: Jatropha bio-diesel production and use. Biomass and Bioenergy, 32(12), 1063-1084. Ackerley, D., B.B.B. Booth, S.H.E. Knight, E.J. Highwood, D.J. Frame, M.R. Allen, and D.P. Rowell, 2011: Sensitivity of Twentieth-Century Sahel Rainfall to Sulfate Aerosol and CO2 Forcing. Journal of Climate, 24(19), 4999-5014. Adam, A.A., M.S. Karsany, and I. Adam, 2010: Manifestations of severe Rift Valley fever in Sudan. International Journal of Infectious Diseases, 14(2), 179-180. ADB, 2011: Climate change and migration in Asia and the Pacific. Asian Development Bank, Metro Manila, Philippines, pp. 77. Adelekan, I.O., 2010: Vulnerability of poor urban coastal communities to flooding in Lagos, Nigeria. Environment and Urbanization, 22(2), 433-450. Adesina, A.A., 2010: Conditioning trends shaping the agricultural and rural landscape in Africa. Agricultural Economics, 41(SUPPL. 1), 73-82. ADF, 2010: Climate change and human development. Proceedings of Seventh African Development Forum (ADF VII) - Acting on climate change for sustainable development in Africa, 10-15 October 2010, Addis Ababa, Ethiopia, pp. 6. Adger, W.N., S. Agrawala, M.M.Q. Mirza, C. Conde, K. O Brien, J. Pulhin, R. Pulwarty, B. Smit, and K. Takahashi, 2007: Assessment of adaptation practices, options, constraints and capacity. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 717-743. Adger, W.N., S. Dessai, M. Goulden, M. Hulme, I. Lorenzoni, D.R. Nelson, L.O. Naess, J. Wolf, and A. Wreford, 2009: Are there social limits to adaptation to climate change? Climatic Change, 93(3-4), 335-354. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Adger, W.N., K. Brown, D.R. Nelson, F. Berkes, H. Eakin, C. Folke, K. Galvin, L. Gunderson, M. Goulden, K. O'Brien, J. Ruitenbeek, and E.L. Tompkins, 2011: Resilience implications of policy responses to climate change. Wiley Interdisciplinary Reviews: Climate Change, 2(5), 757-766. AfDB, ECA, AUC, and UNDP-RBA, 2010: Assessing progress in Africa towards the Millennium Development Goals. MDG Report 2010. African Development Bank (AfDB), Economic Commission for Africa (ECA), African Union Commission (AUC) and United Nations Development Programme Regional Bureau for Africa (UNDP-RBA), pp. 103. AfDB, 2011: The cost of adaptation to climate change in Africa. African Development Bank (AfDB), pp. 41. AfDB, 2012: Congo Basin MRV Regional Project - Phase 1 Multinational: Project Appraisal Report. ADB/BD/WP/2012/47/Approval. African Development Bank (AfDB), pp. 17. Afrane, Y.A., A.K. Githeko, and G. Yan, 2012: The ecology of Anopheles mosquitoes under climate change: Case studies from the effects of deforestation in East African highlands. Annals of the New York Academy of Sciences, 1249(1), 204-210. Agier, L., A. Deroubaix, N. Martiny, P. Yaka, A. Djibo, and H. Broutin, 2013: Seasonality of meningitis in Africa and climate forcing: Aerosols stand out. Journal of the Royal Society Interface, 10(79). Aguilar, E., A.A. Barry, M. Brunet, L. Ekang, A. Fernandes, M. Massoukina, J. Mbah, A. Mhanda, D.J. do Nascimento, T.C. Peterson, O.T. Umba, M. Tomou, and X. Zhang, 2009: Changes in temperature and precipitation extremes in western central Africa, Guinea Conakry, and Zimbabwe, 1955-2006. Journal of Geophysical Research D: Atmospheres, 114(2). Ahmed, S.A., N.S. Diffenbaugh, T.W. Hertel, D.B. Lobell, N. Ramankutty, A.R. Rios, and P. Rowhani, 2011: Climate volatility and poverty vulnerability in Tanzania. Global Environmental Change, 21(1), 46-55. Akullo, D., R. Kanzikwer, P. Birungi, W. Akum, L. Aliguma, and M. Barwogeza, 2007: Indigenous Knowledge in Agriculture: A case study of the challenges in sharing knowledge of past generations in a globalized context in Uganda. Proceedings of World Library and Information Congress: 73rd IFLA General Conference and Council, 19-23, August 2007, Durban, South Africa, pp. 14. Alcamo, J., M. Flörke, and M. Märker, 2007: Future long-term changes in global water resources driven by socio- economic and climatic changes. Hydrological Sciences Journal, 52(2), 247-275. Alem, Y. and M. Söderbom, 2012: Household-Level Consumption in Urban Ethiopia: The Effects of a Large Food Price Shock. World Development, 40(1), 146-162. Al-Gamal, S. and A.K. Dodo, 2009: Impacts of climate changes on water resources in africa with emphasis on groundwater. Journal of Environmental Hydrology, 17, 1-11. Allison, E.H., A.L. Perry, M.-C. Badjeck, W. Neil Adger, K. Brown, D. Conway, A.S. Halls, G.M. Pilling, J.D. Reynolds, N.L. Andrew, and N.K. Dulvy, 2009: Vulnerability of national economies to the impacts of climate change on fisheries. Fish and Fisheries, 10(2), 173-196. Alo, C.A. and G. Wang, 2008: Hydrological impact of the potential future vegetation response to climate changes projected by 8 GCMs. Journal of Geophysical Research G: Biogeosciences, 113(3). Alonso, D., M.J. Bouma, and M. Pascual, 2011: Epidemic malaria and warmer temperatures in recent decades in an East African highland. Proceedings of the Royal Society B: Biological Sciences, 278(1712), 1661-1669. AMCEN and UNEP, 2006: Africa Environment Outlook 2: Our environment, our wealth. United Nations Environment Programme and African Ministerial Conference on the Environment, Nairobi, Kenya, pp. 542. Amigun, B., J.K. Musango, and W. Stafford, 2011: Biofuels and sustainability in Africa. Renewable and Sustainable Energy Reviews, 15(2), 1360-1372. Andriamandimby, S.F., A.E. Randrianarivo-Solofoniaina, E.M. Jeanmaire, L. Ravololomanana, L.T. Razafimanantsoa, T. Rakotojoelinandrasana, J. Razainirina, J. Hoffmann, J.-P. Ravalohery, J.-T. Rafisandratantsoa, P.E. Rollin, and J.-M. Reynes, 2010: Rift valley fever during rainy seasons, Madagascar, 2008 and 2009. Emerging Infectious Diseases, 16(6), 963-970. Anthony, K.R.N., D.I. Kline, G. Diaz-Pulido, S. Dove, and O. Hoegh-Guldberg, 2008: Ocean acidification causes bleaching and productivity loss in coral reef builders. Proceedings of the National Academy of Sciences of the United States of America, 105(45), 17442-17446. Anthony, K.R.N., J.A. Maynard, G. Diaz-Pulido, P.J. Mumby, P.A. Marshall, L. Cao, and O. Hoegh-Guldberg, 2011: Ocean acidification and warming will lower coral reef resilience. Global Change Biology, 17(5), 1798- 1808. Anyah, R.O. and W. Qiu, 2012: Characteristic 20th and 21st century precipitation and temperature patterns and changes over the Greater Horn of Africa. International Journal of Climatology, 32(3), 347-363. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Anyamba, A., J.-P. Chretien, J. Small, C.J. Tucker, P.B. Formenty, J.H. Richardson, S.C. Britch, D.C. Schnabel, R.L. Erickson, and K.J. Linthicum, 2009: Prediction of a Rift Valley fever outbreak. Proceedings of the National Academy of Sciences of the United States of America, 106(3), 955-959. Anyamba, A., K.J. Linthicum, J. Small, S.C. Britch, E. Pak, S. De La Rocque, P. Formenty, A.W. Hightower, R.F. Breiman, J.-P. Chretien, C.J. Tucker, D. Schnabel, R. Sang, K. Haagsma, M. Latham, H.B. Lewandowski, S.O. Magdi, M.A. Mohamed, P.M. Nguku, J.-M. Reynes, and R. Swanepoel, 2010: Prediction, assessment of the Rift Valley fever activity in east and southern Africa 2006-2008 and possible vector control strategies. American Journal of Tropical Medicine and Hygiene, 83(2 SUPPL.), 43-51. Anyamba, A., K.J. Linthicum, J.L. Small, K.M. Collins, C.J. Tucker, E.W. Pak, S.C. Britch, J.R. Eastman, J.E. Pinzon, and K.L. Russell, 2012: Climate teleconnections and recent patterns of human and animal disease outbreaks. PLoS Neglected Tropical Diseases, 6(1), e1465. Aoun, K., F. Amri, E. Chouihi, N. Haouas, K. Bedoui, R. Benikhlef, J. Ghrab, H. Babba, M.K. Chahed, Z. Harrat, and A. Bouratbine, 2008: Epidemiology of Leishmania (L.) infantum. L major and L. killicki in Tunisia: Results and analysis of the identification of 226 human and canine isolates. Bulletin De La Societe De Pathologie Exotique, 101(4), 323-328. Appeaning Addo, K., M. Walkden, and J.P. Mills, 2008: Detection, measurement and prediction of shoreline recession in Accra, Ghana. ISPRS Journal of Photogrammetry and Remote Sensing, 63(5), 543-558. Archer, E.R.M., N.M. Oettlé, R. Louw, and M.A. Tadross, 2008: 'Farming, on the edge' in arid western South Africa: Climate change and agriculture in marginal environments. Geography, 93(2), 98-107. Arendse, A. and T.A. Crane, 2010: Impacts of climate change on smallholder farmers in Africa and their adaptation strategies: what are the roles for research?. International Symposium and Consultation, 29-31 March 2010, CIAT Working Document No. 221. Centro Internacional de Agricultura Tropical (CIAT) and Pan-Africa Bean Research Alliance (PABRA), Kampala, Uganda, pp. 29. Armitage, D. and R. Plummer (eds.), 2010: Adaptive Capacity and Environmental Governance. In: Springer Series on Environmental Management. Springer, Heidelberg, Dordrecht, London, New York, pp. 308. Arndt, C., S. Msangi, and J. Thurlow, 2011: Are biofuels good for African development? An analytical framework with evidence from Mozambique and Tanzania. Biofuels, 2(2), 221-234. Arthurton, R., K. Korateng, T. Forbes, M. Snoussi, J. Kitheka, J. Robinson, N. Shah, S. Taljaard, and P. Monteiro, 2006: Chapter 5: Coastal and marine environments. In: Africa Environment Outlook : Our environment, our wealth. United Nations Environment Programme and African Ministerial Conference on the Environment, Nairobi, Kenya, pp. 155-195. Artur, L. and D. Hilhorst, 2012: Everyday realities of climate change adaptation in Mozambique. Global Environmental Change, 22(2), 529-536. Artzy-Randrup, Y., D. Alonso, and M. Pascual, 2010: Transmission intensity and drug resistance in malaria population dynamics: Implications for climate change. PLoS ONE, 5(10). Atteridge, A., 2011: Will private finance support climate change adaptation in developing countries? Historical investment patterns as a window on future private climate finance. SEI Working Paper No. 2011-05. Stockholm Environment Institute, Stockholm, Sweden, pp. 35. Avery, S., 2012: Lake Turkana and the Lower Omo: Hydrological impacts of major dam and irrigation developments. Volume I. African Studies Centre, University of Oxford, pp. 17. Awuor, C.B., V.A. Orindi, and A. Ochieng Adwera, 2008: Climate change and coastal cities: The case of Mombasa, Kenya. Environment and Urbanization, 20(1), 231-242. Aylett, A., 2010: Conflict, Collaboration and Climate Change: Participatory Democracy and Urban Environmental Struggles in Durban, South Africa. International Journal of Urban and Regional Research, 34(3), 478-495. Azongo, D.K., T. Awine, G. Wak, F.N. Binka, and A.R. Oduro, 2012: A time series analysis of weather variability and all-cause mortality in the Kasena-Nankana districts of Northern Ghana, 1995-2010. Global Health Action, 5(SUPPL.), 14-22. Babiker, S.M., H.D. Blankespoor, M. Wassila, A. Fenwick, and A.A. Daffalla, 1985: Transmission of Schistosoma haematobium in North Gezira, Sudan. Journal of Tropical Medicine and Hygiene, 88(2), 65-73. Badjeck, M.-C., E.H. Allison, A.S. Halls, and N.K. Dulvy, 2010: Impacts of climate variability and change on fishery-based livelihoods. Marine Policy, 34(3), 375-383. Bailey, R., 2013: Managing famine risk: linking early warning to early action. Chatham House, The Royal Institute of International Affairs, London, UK, pp. 100. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bakun, A., D.B. Field, A. Redondo-Rodriguez, and S.J. Weeks, 2010: Greenhouse gas, upwelling-favorable winds, and the future of coastal ocean upwelling ecosystems. Global Change Biology, 16(4), 1213-1228. Banerjee, S., A. Diallo, and Q. Wodon, 2007: Measuring trends in access to modern infrastructure in sub-Saharan Africa: Results from demographic and health surveys. World BankFindings No. 281: Poverty Data, Measurement and Policy Special Expanded Edition. The World Bank. Baoua, I.B., L. Amadou, V. Margam, and L.L. Murdock, 2012: Comparative evaluation of six storage methods for postharvest preservation of cowpea grain. Journal of Stored Products Research, 49, 171-175. Barbier, B., H. Yacouba, H. Karambiri, M. Zoromé, and B. Somé, 2009: Human vulnerability to climate variability in the sahel: Farmers' adaptation strategies in northern burkina faso. Environmental Management, 43(5), 790- 803. Bargaoui, Z., Y. Tramblay, E.A. Lawin, and E. Servat, 2013: Seasonal precipitation variability in regional climate simulations over Northern basins of Tunisia. International Journal of Climatology, Published Online 12-03- 2013. Barkhordarian, A., J. Bhend, and H. von Storch, 2012a: Consistency of observed near surface temperature trends with climate change projections over the Mediterranean region. Climate Dynamics, 38(9-10), 1695-1702. Barkhordarian, A., H. Von Storch, and E. Zorita, 2012b: Anthropogenic forcing is a plausible explanation for the observed surface specific humidity trends over the Mediterranean area. Geophysical Research Letters, 39(19). Barkhordarian, A., H. von Storch, and J. Bhend, 2013: The expectation of future precipitation change over the Mediterranean region is different from what we observe. Climate Dynamics, 40(1-2), 225-244. Barthel, R., B.G.J.S. Sonneveld, J. Götzinger, M.A. Keyzer, S. Pande, A. Printz, and T. Gaiser, 2009: Integrated assessment of groundwater resources in the Ouémé basin, Benin, West Africa. Physics and Chemistry of the Earth, 34(4-5), 236-250. Batisani, N., 2011: Spatio-temporal ephemeral streamflow as influenced by climate variability in Botswana. Journal of Geographical Sciences, 21(3), 417-428. Battersby, J., 2012: Urban food security and climate change: A system of flows. In: Climate Change, Assets and Food Security in Southern African Cities. [Frayne, B., C. Moser, and G. Ziervogel (eds.)]. Earthscan, Abingdon, pp. 35-56. Battisti, D.S. and R.L. Naylor, 2009: Historical warnings of future food insecurity with unprecedented seasonal heat. Science, 323(5911), 240-244. Bauer, S. and I. Scholz, 2010: Adaptation to climate change in Southern Africa: New boundaries for sustainable development? Climate and Development, 2(2), 83-93. Beardon, H. and K. Newman (eds.), 2011: How wide are the ripples? From local participation to international organisational learning. In: Participatory Learning and Action (PLA) No. 63. International Institute for Environment and Development (IIED), London, UK, pp. 186. Beck, L. and T. Bernauer, 2011: How will combined changes in water demand and climate affect water availability in the Zambezi river basin? Global Environmental Change, 21(3), 1061-1072. Beilfuss, R., 2012: A risky climate for Southern African hydro: Assessing hydrological risks and sequences for Zambezi River basin dams. International Rivers, Berkeley, CA, USA, pp. 56. Bele, Y., E. Mulotwa, B. Bokoto de Semboli, D. Sonwa, and A.-M. Tiani, 2010: The Effects of Climate Change in the Congo Basin: The Need to Support Local Adaptive Capacity. Adaptation Insights No. 3: Addressing Climate Change Adaptation in Africa through Participatory Action Research. International Development Research Centre (IDRC) and Department for International Development (DFID), pp. 4. Bele, M.Y., O. Somorin, D.J. Sonwa, J.N. Nkem, and B. Locatelli, 2011: Forests and climate change adaptation policies in Cameroon. Mitigation and Adaptation Strategies for Global Change, 16(3), 369-385. Bellotti, A., B.V.H. Campo, and G. Hyman, 2012: Cassava Production and Pest Management: Present and Potential Threats in a Changing Environment. Tropical Plant Biology, 5(1), 39-72. Below, T.B., K.D. Mutabazi, D. Kirschke, C. Franke, S. Sieber, R. Siebert, and K. Tscherning, 2012: Can farmers' adaptation to climate change be explained by socio-economic household-level variables? Global Environmental Change, 22(1), 223-235. Béné, C., S. Devereux, and R. Sabates-Wheeler, 2012: Shocks and social protection in the Horn of Africa: Analysis from the Productive Safety Net Programme in Ethiopia. CSP Working Paper Number 005 / IDS Working Paper Number 395. Centre for Social Protection (CSP), Institute of Development Studies (IDS), Brighton, UK, pp. 120. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Berchie, J.N., M. Opoku, H. Adu-Dapaah, A. Agyemang, J. Sarkodie-Addo, E. Asare, J. Addo, and H. Akuffo, 2012: Evaluation of five bambara groundnut (Vigna subterranea (L.) Verdc.) landraces to heat and drought stress at Tono-Navrongo, Upper East Region of Ghana. African Journal of Agricultural Research, 7(2), 250- 256. Berenter, J., 2012: Ground Truthing Vulnerability and Adaptation in Africa. [Busby, J.W. (ed.)]. Program on Climate Change and African Political Stability. Robert S. Strauss Center for International Security and Law, Austin, Texas, USA, pp. 17. Berg, A., N. De Noblet-Ducoudré, B. Sultan, M. Lengaigne, and M. Guimberteau, 2013: Projections of climate change impacts on potential C4 crop productivity over tropical regions. Agricultural and Forest Meteorology, 170, 89-102. Bergengren, J.C., D.E. Waliser, and Y.L. Yung, 2011: Ecological sensitivity: A biospheric view of climate change. Climatic Change, 107(3), 433-457. Berkes, F., 2009: Evolution of co-management: Role of knowledge generation, bridging organizations and social learning. Journal of Environmental Management, 90(5), 1692-1702. Berrang-Ford, L., J.D. Ford, and J. Paterson, 2011: Are we adapting to climate change? Global Environmental Change, 21(1), 25-33. Beyene, T., D.P. Lettenmaier, and P. Kabat, 2010: Hydrologic impacts of climate change on the Nile River Basin: Implications of the 2007 IPCC scenarios. Climatic Change, 100(3), 433-461. Biasutti, M. and A. Giannini, 2006: Robust Sahel drying in response to late 20th century forcings. Geophysical Research Letters, 33(11). Biasutti, M., I.M. Held, A.H. Sobel, and A. Giannini, 2008: SST forcings and Sahel rainfall variability in simulations of the twentieth and twenty-first centuries. Journal of Climate, 21(14), 3471-3486. Biasutti, M., 2013: Forced Sahel rainfall trends in the CMIP5 archive. Journal of Geophysical Research D: Atmospheres, 118(4), 1613-1623. Biazin, B., G. Sterk, M. Temesgen, A. Abdulkedir, and L. Stroosnijder, 2012: Rainwater harvesting and management in rainfed agricultural systems in sub-Saharan Africa - A review. Physics and Chemistry of the Earth, 47-48, 139-151. Biesbroek, G.R., R.J. Swart, and W.G.M. van der Knaap, 2009: The mitigation-adaptation dichotomy and the role of spatial planning. Habitat International, 33(3), 230-237. Bizikova, L., S. Boardley, and S. Mead, 2010: Participatory scenario development approaches for identifying pro- poor adaptation options. In: Economics of adaptation to climate change. Discussion Paper Number 18, December 2010. The World Bank, Washington, DC, USA, pp. 67. Black, R., W.N. Adger, N.W. Arnell, S. Dercon, A. Geddes, and D. Thomas, 2011a: The effect of environmental change on human migration. Global Environmental Change, 21(SUPPL. 1), S3-S11. Black, R., S.R.G. Bennett, S.M. Thomas, and J.R. Beddington, 2011b: Climate change: Migration as adaptation. Nature, 478(7370), 447-449. Black, R., D. Kniveton, and K. Schmidt-Verkerk, 2011c: Migration and climate change: Towards an integrated assessment of sensitivity. Environment and Planning A, 43(2), 431-450. Blanford, J.I., S. Blanford, R.G. Crane, M.E. Mann, K.P. Paaijmans, K.V. Schreiber, and M.B. Thomas, 2013: Implications of temperature variation for malaria parasite development across Africa. Scientific Reports, 3, 1300. BNRCC, 2012: Towards a Lagos State Climate Change Adaptation Strategy. Report prepared for the Commissioner of Environment, Lagos State, January 2012. Building Nigeria's Response to Climate Change (BNRCC) project, pp. 46. Boateng, I., 2006: Shoreline management planning: can it benefit Ghana? A case study of UK SMPs and their potential relevance in Ghana. Proceedings of 5th FIG Regional Conference: Promoting Land Administration and Good Governance, 8-11, March 2006, Accra, pp. 17. Boko, M., I. Niang, A. Nyong, C. Vogel, A. Githeko, M. Medany, B. Osman-Elasha, R. Tabo, and P. Yanda, 2007: Africa. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 433-467. Bomblies, A. and E.A.B. Eltahir, 2010: Assessment of the Impact of Climate Shifts on Malaria Transmission in the Sahel. EcoHealth, 6(3), 426-437. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bond, W.J. and G.F. Midgley, 2012: Carbon dioxide and the uneasy interactions of trees and savannah grasses. Philosophical Transactions of the Royal Society B: Biological Sciences, 367(1588), 601-612. Born, K., M. Christoph, A.H. Fink, P. Knippertz, H. Paeth, and P. Speth, 2008: Moroccan Climate in the Present and Future: Combined View from Observational Data and Regional Climate Scenarios. In: Climatic Changes and Water Resources in the Middle East and North Africa. [Zereini, F. and H. Hötzl (eds.)]. Springer, Berlin, Heidelberg, pp. 29-45. Bouchaou, L., J.L. Michelot, A. Vengosh, Y. Hsissou, M. Qurtobi, C.B. Gaye, T.D. Bullen, and G.M. Zuppi, 2008: Application of multiple isotopic and geochemical tracers for investigation of recharge, salinization, and residence time of water in the Souss-Massa aquifer, southwest of Morocco. Journal of Hydrology, 352(3-4), 267-287. Bouma, M.J., A. Baeza, A. terVeen, and M. Pascual, 2011: Global malaria maps and climate change: A focus on East African highlands. Trends in Parasitology, 27(10), 421-422. Bounoua, L., K. Kahime, L. Houti, T. Blakey, K.L. Ebi, P. Zhang, M.L. Imhoff, K.J. Thome, C. Dudek, S.A. Sahabi, M. Messouli, B. Makhlouf, A.E. Laamrani, and A. Boumezzough, 2013: Linking Climate to Incidence of Zoonotic Cutaneous Leishmaniasis (L. major) in Pre-Saharan North Africa. International Journal of Environmental Research and Public Health, 10(8), 3172-3191. Brasseur, G.P., M. Schultz, C. Granier, M. Saunois, T. Diehl, M. Botzet, E. Roeckner, and S. Walters, 2006: Impact of climate change on the future chemical composition of the global troposphere. Journal of Climate, 19(16), 3932-3951. Braune, E. and Y. Xu, 2010: The role of ground water in sub-Saharan Africa. Ground Water, 48(2), 229-238. Brink, A.B. and H.D. Eva, 2009: Monitoring 25 years of land cover change dynamics in Africa: A sample based remote sensing approach. Applied Geography, 29(4), 501-512. Brockhaus, M. and H. Djoudi, 2008: Adaptation at the interface of forest ecosystem goods and services and livestock production systems in Northern Mali. CIFOR infobriefs No. 19. Center for International Forestry Research (CIFOR), Bogor, Indonesia, pp. 4. Brooks, N., W.N. Adger, and P.M. Kelly, 2005: The determinants of vulnerability and adaptive capacity at the national level and the implications for adaptation. Global Environmental Change, 15(2), 151-163. Brooks, N., N. Grist, and K. Brown, 2009: Development futures in the context of climate change: Challenging the present and learning from the past. Development Policy Review, 27(6), 741-765. Brown, T., S. Gibson, and S. Ashley, 2007: Building consensus for social protection: Insights from Ethiopia's Productive Safety Net Programme (PSNP). The IDL Group, pp. 10. Brown, D.R., P. Dettmann, T. Rinaudo, H. Tefera, and A. Tofu, 2011: Poverty alleviation and environmental restoration using the clean development mechanism: A case study from Humbo, Ethiopia. Environmental Management, 48(2), 322-333. Brown, H.C.P., J.N. Nkem, D.J. Sonwa, and Y. Bele, 2010: Institutional adaptive capacity and climate change response in the Congo Basin forests of Cameroon. Mitigation and Adaptation Strategies for Global Change, 15(3), 263-282. Brown, M.E., B. Hintermann, and N. Higgins, 2009: Markets, climate change, and food security in West Africa. Environmental Science and Technology, 43(21), 8016-8020. Bryan, E., C. Ringler, B. Okoba, C. Roncoli, S. Silvestri, and M. Herrero, 2011: Adapting agriculture to climate change in Kenya: Household and community strategies and determinants. pp. 30. Bryan, E., T.T. Deressa, G.A. Gbetibouo, and C. Ringler, 2009: Adaptation to climate change in Ethiopia and South Africa: options and constraints. Environmental Science and Policy, 12(4), 413-426. Buchner, B., A. Falconer, M. Hervé-Mignucci, and C. Trabacchi, 2012: The landscape of climate finance 2012. A CPI Report. Climate Policy Initiative, San Francisco, CA, USA, pp. 84. Buhaug, H., 2010: Climate not to blame for African civil wars. Proceedings of the National Academy of Sciences of the United States of America, 107(38), 16477-16482. Buitenwerf, R., W.J. Bond, N. Stevens, and W.S.W. Trollope, 2012: Increased tree densities in South African savannas: >50 years of data suggests CO2 as a driver. Global Change Biology, 18(2), 675-684. Bunce, M., K. Brown, and S. Rosendo, 2010a: Policy misfits, climate change and cross-scale vulnerability in coastal Africa: how development projects undermine resilience. Environmental Science and Policy, 13(6), 485-497. Bunce, M., S. Rosendo, and K. Brown, 2010b: Perceptions of climate change, multiple stressors and livelihoods on marginal African coasts. Environment, Development and Sustainability, 12(3), 407-440. Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Burke, M.B., D.B. Lobell, and L. Guarino, 2009a: Shifts in African crop climates by 2050, and the implications for crop improvement and genetic resources conservation. Global Environmental Change, 19(3), 317-325. Burke, M.B., E. Miguel, S. Satyanath, J.A. Dykema, and D.B. Lobell, 2009b: Warming increases the risk of civil war in Africa. Proceedings of the National Academy of Sciences of the United States of America, 106(49), 20670-20674. Burney, J.A. and R.L. Naylor, 2012: Smallholder Irrigation as a Poverty Alleviation Tool in Sub-Saharan Africa. World Development, 40(1), 110-123. Burton, I., S. Huq, B. Lim, O. Pilifosova, and E.L. Schipper, 2002: From impacts assessment to adaptation priorities: The shaping of adaptation policy. Climate Policy, 2(2-3), 145-159. Busby, J.W., T.G. Smith, K.L. White, and S.M. Strange, 2013: Climate change and insecurity: Mapping vulnerability in Africa. International Security, 37(4), 132-172. Bwango, A., J. Wright, C. Elias, and I. Burton, 2000: Reconciling national and global priorities in adaptation to climate change: With an illustration from Uganda. Environmental Monitoring and Assessment, 61(1), 145-159. Calow, R. and A. MacDonald, 2009: What will climate change mean for groundwater supply in Africa?. ODI Background Notes. Overseas Development Institute, London, UK, pp. 8. Caminade, C., J.A. Ndione, C.M.F. Kebe, A.E. Jones, S. Danuor, S. Tay, Y.M. Tourre, J.-P. Lacaux, C. Vignolles, J.B. Duchemin, I. Jeanne, and A.P. Morse, 2011: Mapping Rift Valley fever and malaria risk over West Africa using climatic indicators. Atmospheric Science Letters, 12(1), 96-103. CARE, 2012: Decision-making for climate resilient livelihoods and risk reduction: a participatory scenario planning approach. CARE International - Adaptation Learning Programme for Africa, pp. 12. CARE International, 2012: Community Based Adaptation - Experiences from Africa. In: jotoafrika - Adapting to Climate Change in Africa [Percy, F. (ed.)]. Special Issue 11. CARE International, Nairobi, Kenya, pp. 8. Carneiro, J.F., M. Boughriba, A. Correia, Y. Zarhloule, A. Rimi, and B.E. Houadi, 2010: Evaluation of climate change effects in a coastal aquifer in Morocco using a density-dependent numerical model. Environmental Earth Sciences, 61(2), 241-252. Carter, R.C. and A. Parker, 2009: Climate change, population trends and groundwater in Africa. Hydrological Sciences Journal, 54(4), 676-689. Cartwright, A., 2008: A sea-level rise risk assessment for the city of Cape Town. Phase three: final report. Stockholm Environment Institute (SEI) and LaquaR Consultants CC, Cape Town, South Africa, pp. 65. Cartwright, A., G. Brundrit, and L. Fairhurst, 2008: A sea-level rise risk assessment for the City of Cape Town. Phase 4. Global Climate Change and Adaptation. Sea-Level Rise Adaptation and risk mitigation measures for the City of Cape Town. LaquaR Consultants CC, Cape Town, South Africa, pp. 42. Castán Broto, V., B. Oballa, and P. Junior, 2013: Governing climate change for a just city: Challenges and lessons from Maputo, Mozambique. Local Environment, 18(6), 678-704. Castro, A.P., D. Taylor, and D.W. Brokensha (eds.), 2012: Climate Change and Threatened Communities. Vulnerability, capacity and action. Practical Action, pp. 224. CCAA, 2011: Extending our reach: Climate Change Adaptation in Africa program annual report 2010-11. Climate Change Adaptation in Africa program, pp. 8. Chambwera, M. and S. Anderson, 2011: Integrating climate change into agricultural research for development in Africa. IIED Briefing Paper. International Institute for Environment and Development (IIED), London, UK, pp. 4. Chandani, A., 2011: Improving leadership on climate change. IIED Reflect & Act series, March 2011. International Institute for Environment and Development (IIED) Climate Change Group, pp. 2. Chang a, L.B., P.Z. Yanda, and J. Ngana, 2010: Indigenous knowledge in seasonal rainfall prediction in Tanzania: A case of the South-western Highland of Tanzania . Journal of Geography and Regional Planning, 3(4), 66-72. Chaves, L.F., A. Kaneko, G. Taleo, M. Pascual, and M.L. Wilson, 2008: Malaria transmission pattern resilience to climatic variability is mediated by insecticide-treated nets. Malaria Journal, 7, 100. Chaves, L.F. and C.J.M. Koenraadt, 2010: Climate change and highland malaria: Fresh air for a hot debate. Quarterly Review of Biology, 85(1), 27-55. Chaves, L.F., M. Hashizume, A. Satake, and N. Minakawa, 2012: Regime shifts and heterogeneous trends in malaria time series from Western Kenya Highlands. Parasitology, 139(1), 14-25. Chevalier, V., T. Rakotondrafara, M. Jourdan, J.M. Heraud, H.R. Andriamanivo, B. Durand, J. Ravaomanana, P.E. Rollin, and R. Rakotondravao, 2011: An unexpected recurrent transmission of rift valley fever virus in cattle in a temperate and mountainous area of madagascar. PLoS Neglected Tropical Diseases, 5(12). Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chevallier, R., 2012: Political barriers to climate change adaptation implementation in SADC. In: Overcoming Barriers to Climate Change Adaptation Implementation in Southern Africa. [Masters, L. and L. Duff (eds.)]. Africa Institute of South Africa, Pretoria, South Africa, pp. 1-19. Chigwada, J., 2005: Case study 6: Zimbabwe. Climate proofing infrastructure and diversifying livelihoods in Zimbabwe. IDS Bulletin, 36(4), 103-116. Chikapa, C., 2012: Impact of radio, cinema and vernacular publications in enhancing community based adaptation to climate change. Proceedings of Sixth International Conference on Community-Based Adaptation, 16-22, April 2012, Hanoi, Vietnam, pp. 64-64. Chikozho, C., 2010: Applied social research and action priorities for adaptation to climate change and rainfall variability in the rainfed agricultural sector of Zimbabwe. Physics and Chemistry of the Earth, 35(13-14), 780- 790. Chishakwe, N., L. Murray, and M. Chambwera, 2012: Building climate change adaptation on community experiences: Lessons from community-based natural resource management in southern Africa. International Institute for Environment and Development (IIED), London, UK, pp. 126. Christensen, J.H., B. Hewitson, A. Busuioc, A. Chen, X. Gao, I. Held, R. Jones, R.K. Kolli, W.-T. Kwon, R. Laprise, V. Magana Rueda, L. Mearns, C.G. Menéndez, J. Räisänen, A. Rinke, A. Sarr, and P. Whetton, 2007: Regional climate projections. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 847-940. Christy, J.R., W.B. Norris, and R.T. McNider, 2009: Surface temperature variations in east Africa and possible causes. Journal of Climate, 22(12), 3342-3356. CIFOR, 2011: Climate change and forests in the Congo Basin: Synergy between adaptation and mitigation. COBAM Project Information number 1. Center for International Forestry Research, pp. 2. Clements, R., J. Haggar, A. Quezada, and J. Torres, 2011: Technologies for climate change adaptation: agriculture sector. In: TNA Guidebook Series [Zhu, X. (ed.)]. UNEP Ris Centre, Roskilde, Denmark, pp. 198. Climate Investment Fund, 2009: Programming and financing modalities for the SCF targeted program, the Pilot Program for Climate Change Resilience (PPCR). Climate Investment Funds, pp. 27. Clover, J. and S. Eriksen, 2009: The effects of land tenure change on sustainability: human security and environmental change in southern African savannas. Environmental Science and Policy, 12(1), 53-70. Codjoe, S.N.A. and G. Owusu, 2011: Climate change/variability and food systems: evidence from the Afram Plains, Ghana. Regional Environmental Change, 11(4), 753-765. Codjoe, S.N.A., L.K. Atidoh, and V. Burkett, 2012: Gender and occupational perspectives on adaptation to climate extremes in the Afram Plains of Ghana. Climatic Change, 110(1-2), 431-454. Coe, R. and R.D. Stern, 2011: Assessing and addressing climate-induced risk in sub-Saharan rainfed agriculture: Lessons learned. Experimental Agriculture, 47(2), 395-410. Cohen, M.J. and J.L. Garrett, 2010: The food price crisis and urban food (in)security. Environment and Urbanization, 22(2), 467-482. Collier, P., G. Conway, and T. Venables, 2008: Climate change and Africa. Oxford Review of Economic Policy, 24(2), 337-353. Collins, J.M., 2011: Temperature variability over Africa. Journal of Climate, 24(14), 3649-3666. Conway, G., 2009: The science of climate change in Africa: impacts and adaptation. Discussion paper No. 1. Grantham Institute for Climate Change, London, UK, pp. 24. Conway, D., 2011: Adapting climate research for development in Africa. Wiley Interdisciplinary Reviews: Climate Change, 2(3), 428-450. Conway, D. and E.L.F. Schipper, 2011: Adaptation to climate change in Africa: Challenges and opportunities identified from Ethiopia. Global Environmental Change, 21(1), 227-237. Cook, K.H. and E.K. Vizy, 2012: Impact of climate change on mid-twenty-first century growing seasons in Africa. Climate Dynamics, 39(12), 2937-2955. Cook, K.H. and E.K. Vizy, 2013: Projected changes in east african rainy seasons. Journal of Climate, 26(16), 5931- 5948. Cooper, P.J.M., J. Dimes, K.P.C. Rao, B. Shapiro, B. Shiferaw, and S. Twomlow, 2008: Coping better with current climatic variability in the rain-fed farming systems of sub-Saharan Africa: An essential first step in adapting to future climate change? Agriculture, Ecosystems and Environment, 126(1-2), 24-35. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Corsi, M., S. Hagemann, and C. Salgado Silva, 2012: Africa Adaptation Programme third quarterly report 2012. Prepared by the AAP Inter-Regional Technical Support Component. UNDP Africa Adaptation Programme, pp. 54. Cotter, M., R. de la Pena-Lavander, and J. Sauerborn, 2012: Understanding the present distribution of the parasitic weed Striga hermonthica and predicting its potential future geographic distribution in the light of climate change. Julius-Kühn-Archiv, 2(434), 630-636. Cotty, P.J. and R. Jaime-Garcia, 2007: Influences of climate on aflatoxin producing fungi and aflatoxin contamination. International Journal of Food Microbiology, 119(1-2), 109-115. Crane, T.A., 2010: Of models and meanings: Cultural resilience in social-ecological systems. Ecology and Society, 15(4), 19. Crawford, R.J.M. and R. Altwegg, 2009: Seabirds and climate change in southern African some considerations[Harebottle, D.M., A.J.F.K. Craig, M.D. Anderson, H. Rakotomanana, M. Muchai(eds.)]. Proceedings of The 12th Pan-African Ornithological Congress, Cape Town, South Africa, pp. 5. Crétat, J., B. Pohl, Y. Richard, and P. Drobinski, 2012: Uncertainties in simulating regional climate of Southern Africa: Sensitivity to physical parameterizations using WRF. Climate Dynamics, 38(3-4), 613-634. Crush, J. and B. Frayne, 2010: Pathways to Insecurity: Urban Food Supply and Access in Southern African Cities. Urban Food Security Series No. 3. African Food Security Urban Network (AFSUN), pp. 46. Cuevas, L.E., I. Jeanne, A. Molesworth, M. Bell, E.C. Savory, S.J. Connor, and M.C. Thomson, 2007: Risk mapping and early warning systems for the control of meningitis in Africa. Vaccine, 25(SUPPL. 1), A12-A17. Dabi, D.D., A.O. Nyong, A.A. Adepetu, and V.I. Ihemegbulem, 2008: Past, present and future adaptation by rural households of northern Nigeria. In: Climate change and adaptation. [Leary, N., J. Adejuwon, V. Barros, I. Burton, J. Kulkarni, and R. Lasco (eds.)]. Earthscan, London, UK and Sterling, VA, USA, pp. 147-162. Dapi, L.N., J. Rocklöv, G. Nguefack-Tsague, E. Tetanye, and T. Kjellstrom, 2010: Heat impact on schoolchildren in Cameroon, Africa: potential health threat from climate change. Global Health Action, 3(5610), 8. Darwall, W., K. Smith, D. Allen, R. Holland, I. Harrison, and E. Brooks (eds.), 2011: The Diversity of life in African freshwaters: under water, under threat. An analysis of the status and distribution of freshwater species throughout mainland Africa IUCN, Gland, Switzerland and Cambridge, UK, pp. 348. Davies, M., B. Guenther, J. Leavy, T. Mitchell, and T. Tanner, 2009: Climate change adaptation, disaster risk reduction and social protection: complementary roles in agriculture rural growth?. IDS Working Paper No. 320, February 2009. Institute of Development Studies (IDS), Brighton, UK, pp. 37. de Magny, G.C., J.-F. Guégan, M. Petit, and B. Cazelles, 2007: Regional-scale climate-variability synchrony of cholera epidemics in West Africa. BMC Infectious Diseases, 7. de Magny, G.C., W. Thiaw, V. Kumar, N.M. Manga, B.M. Diop, L. Gueye, M. Kamara, B. Roche, R. Murtugudde, and R.R. Colwell, 2012: Cholera Outbreak in Senegal in 2005: Was Climate a Factor? PLoS ONE, 7(8). De Silva, S.S. and D. Soto, 2009: Climate change and aquaculture: potential impacts, adaptation and mitigation. In: Climate change implications for fisheries and aquaculture: overview of current scientific knowledge. FAO Fisheries and Aquaculture Technical Paper No. 530. [Cochrane, K., C. De Young, D. Soto, and T. Bahri (eds.)]. Food and Agriculture Organization of the United Nations, Rome, Italy, pp. 151-212. De Wit, M. and J. Stankiewicz, 2006: Changes in surface water supply across Africa with predicted climate change. Science, 311(5769), 1917-1921. Debsu, D.N., 2012: Community based coping mechanisms and adaptations to droughts in the Borana Pastoral area of southern Ethiopia. Proceedings of Sixth International Conference on Community-Based Adaptation, 16-22, April 2012, Hanoi, Vietnam, pp. 42-42. Delire, C., A. Ngomanda, and D. Jolly, 2008: Possible impacts of 21st century climate on vegetation in Central and West Africa. Global and Planetary Change, 64(1-2), 3-15. Denman, K.L., G. Brasseur, A. Chidthaisong, P. Ciais, P.M. Cox, R.E. Dickinson, D. Hauglustaine, C. Heinze, E. Holland, D. Jacob, U. Lohmann, S. Ramachandran, P.L. da Silva Dias, S.C. Wofsy, and X. Zhang, 2007: Couplings Between Changes in the Climate System and Biogeochemistry. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 499-587. Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Denton, F., S. Anderson, and J. Ayers, 2011: Getting African climate change research recognised. IIED Briefing, Lessons from Adaptation in Practice, November 2011. International Institute for Environment and Development (IIED), London, UK, pp. 4. Denton, F., 2010: Financing adaptation in Least Developed Countries in West Africa: Is finance the 'real deal'? Climate Policy, 10(6), 655-671. Deressa, T.T., R.M. Hassan, C. Ringler, T. Alemu, and M. Yesuf, 2009: Determinants of farmers' choice of adaptation methods to climate change in the Nile Basin of Ethiopia. Global Environmental Change, 19(2), 248- 255. Desanker, P., C. Magdaza, A. Allalli, C. Basalirwa, M. Boko, G. Dieudonne, T.E. Downing, P.O. Dube, A. Githeko, M. Githendu, P. Gonzalez, D. Gwary, B. Jallow, J. Nwafa, and R. Scholes, 2001: Africa. In: Climate Change 2001. Impacts, Adaptation, and Vulnerability. [Mc Carthy, J.J., O.F. Canziani, N.A. Leary, and D.J. Dokken (eds.)]. Cambridge University Press, Cambridge, UK, pp. 487-531. Descheemaeker, K., T. Amede, A. Haileslassie, and D. Bossio, 2011: Analysis of gaps and possible interventions for improving water productivity in crop livestock systems of Ethiopia. Experimental Agriculture, 47(S1), 21-38. Descheemaeker, K., E. Mapedza, T. Amede, and W. Ayalneh, 2010: Effects of integrated watershed management on livestock water productivity in water scarce areas in Ethiopia. Physics and Chemistry of the Earth, 35(13-14), 723-729. Dessu, S.B. and A.M. Melesse, 2012: Impact and uncertainties of climate change on the hydrology of the Mara River basin, Kenya/Tanzania. Hydrological Processes, 27(20), 2973-2986. Devitt, C. and R.S.J. Tol, 2012: Civil war, climate change, and development: A scenario study for sub-saharan Africa. Journal of Peace Research, 49(1), 129-145. Diagne, K., 2007: Governance and natural disasters: Addressing flooding in Saint Louis, Senegal. Environment and Urbanization, 19(2), 552-562. Diallo, I., M.B. Sylla, F. Giorgi, A.T. Gaye, and M. Camara, 2012: Multimodel GCM-RCM ensemble-based projections of temperature and precipitation over West Africa for the Early 21st Century. International Journal of Geophysics, 2012. Diboulo, E., A. Sié, J. Rocklöv, L. Niamba, M. Yé, C. Bagagnan, and R. Sauerborn, 2012: Weather and mortality: A 10 year retrospective analysis of the nouna health and demographic surveillance system, Burkina Faso. Global Health Action, 5(SUPPL.), 6-13. Diffenbaugh, N.S. and F. Giorgi, 2012: Climate change hotspots in the CMIP5 global climate model ensemble. Climatic Change, 114(3-4), 813-822. Dillon, A., 2011: The effect of irrigation on poverty reduction, asset accumulation, and informal insurance: Evidence from Northern Mali. World Development, 39(12), 2165-2175. Dinku, T., T. Arivelo, S.B. Awulachew, A.F. Kamgaf, S.A. Moges, B.S. Nyenzi, and Y. Sileshi, 2011: Climate science, information, and services in Africa: Status, gaps and policy implications. Working Paper 1, November 2011. United Nations Economic Commission for Africa, African Climate Policy Centre, pp. 26. Diop, S., R. Arthurton, P. Scheren, J. Kitheka, K. Koranteng, and R. Payet, 2011: The coastal and marine environment of Western and Eastern Africa: challenges to sustainable management and socioeconomic development. In: Treatise on Estuarine and Coastal Science. Academic Press, Amsterdam, pp. 315-335. Djoudi, H. and M. Brockhaus, 2011: Is adaptation to climate change gender neutral? Lessons from communities dependent on livestock and forests in northern Mali. International Forestry Review, 13(2), 123-135. Djoudi, H., M. Brockhaus, and B. Locatelli, 2013: Once there was a lake: Vulnerability to environmental changes in northern Mali. Regional Environmental Change, 13(3), 493-508. Dodman, D. and J. Carmin, 2011: Urban adaptation planning: the use and limits of climate science. IIED Briefing: Lessons from adaptation in practice. International Institute for Environment and Development (IIED), London, UK, pp. 4. Dondji, B., 2001: Leishmaniasis and phlebotomine sandflies of Cameroon: A review. Bulletin De La Societe De Pathologie Exotique, 94(3), 277-279. Dondorp, A.M., S. Yeung, L. White, C. Nguon, N.P.J. Day, D. Socheat, and L. Von Seidlein, 2010: Artemisinin resistance: Current status and scenarios for containment. Nature Reviews Microbiology, 8(4), 272-280. Dossou, K.M.R. and B. Gléhouenou-Dossou, 2007: The vulnerability to climate change of Cotonou (Benin): The rise in sea level. Environment and Urbanization, 19(1), 65-79. Dougill, A.J., E.D.G. Fraser, and M.S. Reed, 2010: Anticipating vulnerability to climate change in dryland pastoral systems: Using dynamic systems models for the Kalahari. Ecology and Society, 15(2), 17. Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Douglas, I., K. Alam, M. Maghenda, Y. Mcdonnell, L. Mclean, and J. Campbell, 2008: Unjust waters: Climate change, flooding and the urban poor in Africa. Environment and Urbanization, 20(1), 187-205. Drimie, S. and S. Gillespie, 2010: Adaptation to climate change in Southern Africa: Factoring in AIDS. Environmental Science and Policy, 13(8), 778-784. Drine, I., 2011: Climate change compounding risks in North Africa. UNU-WIDER Working Paper No. 2011/32. United Nations University, World Institute for Development Economics Research, pp. 19. Driouech, F., M. Déqué, and E. Sánchez-Gómez, 2010: Weather regimes-Moroccan precipitation link in a regional climate change simulation. Global and Planetary Change, 72(1-2), 1-10. Droogers, P., W.W. Immerzeel, W. Terink, J. Hoogeveen, M.F.P. Bierkens, L.P.H. Van Beek, and B. Debele, 2012: Water resources trends in Middle East and North Africa towards 2050. Hydrology and Earth System Sciences, 16(9), 3101-3114. Druyan, L.M., 2011: Studies of 21st-century precipitation trends over West Africa. International Journal of Climatology, 31(10), 1415-1424. Dube, O.P. and M.B.M. Sekhwela, 2007: Community Coping Strategies in Semiarid Limpopo Basin Part of Botswana: Enhancing Adaptation Capacity to Climate Change. AIACC Working Paper No. 47. International START Secretariat, Washington, DC, pp. 42. Dube, O.P. and B.M. Sekhwela, 2008: Indigenous knowledge, institutions and practices for coping with variable climate in the Limpopo Basin of Botswana. In: Climate change and adaptation. [Leary, N., J. Adejuwon, V. Barros, I. Burton, J. Kulkarni, and R. Lasco (eds.)]. Earthscan, London, UK and Sterling, VA, USA, pp. 71-89. Dukiæ, V., M. Hayden, A.A. Forgor, T. Hopson, P. Akweongo, A. Hodgson, A. Monaghan, C. Wiedinmyer, T. Yoksas, M.C. Thomson, S. Trzaska, and R. Pandya, 2012: The Role of Weather in Meningitis Outbreaks in Navrongo, Ghana: A Generalized Additive Modeling Approach. Journal of Agricultural, Biological, and Environmental Statistics, 17(3), 442-460. Easterly, W., 2009: Can the west save Africa? Journal of Economic Literature, 47(2), 373-447. Ebi, K.L., J. Padgham, M. Doumbia, A. Kergna, J. Smith, T. Butt, and B. McCarl, 2011: Smallholders adaptation to climate change in Mali. Climatic Change, 108(3), 423-436. ECA, AUC, and AfDB, 2009: Assessing progress in Africa toward the Millennium Development Goals. MDG Report 2009. Economic Commission for Africa (ECA), African Union Commission (AUC) and African Development Bank Group (AfDB), Addis Ababa, Ethiopia, pp. 70. ECA and AUC, 2011: Economic Report on Africa 2011: Governing development in Africa - the role of the state in economic transformation. United Nations Economic Commission for Africa (ECA) and African Union Commission (AUC), Addis Ababa, Ethiopia, pp. 130. ECA, AUC, AfDB, and UNDP RBA, 2012: Assessing progress in Africa toward the Millennium Development Goals. Emerging perspectives from Africa on the post-2015 development agenda. MDG Report 2012. Economic Commission for Africa (ECA), African Union Commission (AUC), African Development Bank (AfDB) and United Nations Development Programme Regional Bureau for Africa (UNDP RBA), pp. 160. Edlund, S., M. Davis, J.V. Douglas, A. Kershenbaum, N. Waraporn, J. Lessler, and J.H. Kaufman, 2012: A global model of malaria climate sensitivity: Comparing malaria response to historic climate data based on simulation and officially reported malaria incidence. Malaria Journal, 11, 331. Egbendewe-Mondzozo, A., M. Musumba, B.A. McCarl, and X. Wu, 2011: Climate change and vector-borne diseases: An economic impact analysis of malaria in Africa. International Journal of Environmental Research and Public Health, 8(3), 913-930. Egondi, T., C. Kyobutungi, S. Kovats, K. Muindi, R. Ettarh, and J. Rocklöv, 2012: Time-series analysis of weather and mortality patterns in Nairobi's informal settlements. Global Health Action, 5(SUPPL.), 23-32. Eid, H.M., S.M. El-Marsafawy, and S.A. Ouda, 2007: Assessing the Economic Impacts of Climate Change on Agriculture in Egypt: A Ricardian Approach. Policy Research Working Paper 4342. The World Bank, pp. 39. Eitzinger, A., P. Läderach, A. Quiroga, A. Pantoja, and J. Gordon, 2011a: Future climate scenarios for Kenya s tea growing areas. Centro Internacional de Agricultura Tropical (CIAT), Cali, Colombia, pp. 27. Eitzinger, A., P. Läderach, A. Quiroga, A. Pantoja, and J. Gordon, 2011b: Future climate scenarios for Uganda s tea growing areas. Centro Internacional de Agricultura Tropical (CIAT), Cali, Colombia, pp. 23. Ellison, J.C. and I. Zouh, 2012: Vulnerability to Climate Change of Mangroves: Assessment from Cameroon, Central Africa. Biology, 1(3), 617-638. Elshamy, M.E., I.A. Seierstad, and A. Sorteberg, 2009: Impacts of climate change on Blue Nile flows using bias- corrected GCM scenarios. Hydrology and Earth System Sciences, 13(5), 551-565. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Engelbrecht, F.A., J.L. McGregor, and C.J. Engelbrecht, 2009: Dynamics of the conformal-cubic atmospheric model projected climate-change signal over southern Africa. International Journal of Climatology, 29(7), 1013-1033. Ensor, J. and R. Berger, 2009: Community-based adaptation and culture in theory and practice. In: Adapting to Climate Change: Thresholds, Values, Governance. [Adger, W.N., I. Lorenzoni, and K. O Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 227-239. Eriksen, S. and J.A. Silva, 2009: The vulnerability context of a savanna area in Mozambique: household drought coping strategies and responses to economic change. Environmental Science and Policy, 12(1), 33-52. Ermert, V., A.H. Fink, A.P. Morse, and H. Paeth, 2012: The impact of regional climate change on malaria risk due to greenhouse forcing and land-use changes in tropical Africa. Environmental Health Perspectives, 120(1), 77- 84. Evans, J.P., 2011: Resilience, ecology and adaptation in the experimental city. Transactions of the Institute of British Geographers, 36(2), 223-237. Eyong, C.T., 2007: Indigenous knowledge and sustainable development in Africa: case study on central Africa. In: Indigenous knowledge systems and sustainable development: relevance for Africa. Tribes and Tribals, Special Volume No. 1. [Boon, E.K. and L. Hens (eds.)]. Kamla-Raj Enterprises, Dheli, India, pp. 121-139. Fankhauser, S. and I. Burton, 2011: Spending adaptation money wisely. Working Paper No. 47 and No. 37. Centre for Climate Change Economics and Policy and Grantham Research Institute on Climate Change and the Environment, London, UK, pp. 22. Fankhauser, S. and G. Schmidt-Traub, 2010: From adaptation to climate-resilient development: the costs of climate- proofing the Millennium Development Goals in Africa. Centre for Climate Change Economics and Policy (CCCEP) and Grantham Research Institute on Climate Change and the Environment, London, UK, pp. 26. Fankhauser, S., 1997: The social costs of climate change: The IPCC second assessment report and beyond. Mitigation and Adaptation Strategies for Global Change, 1(4), 385-403. FAO, 2002: World agriculture: towards 2015/2030. Summary Report. Food and Agriculture Organization of the United Nations, Rome, Italy, pp. 97. FAO, 2011: Disaster Risk Management Strategy in West Africa and the Sahel. Food and Agriculture Organization of the United Nations (FAO), Italy, Rome, pp. 41. Fatondji, D., C. Martius, R. Zougmore, P.L.G. Vlek, C.L. Bielders, and S. Koala, 2009: Decomposition of organic amendment and nutrient release under the zai technique in the Sahel. Nutrient Cycling in Agroecosystems, 85(3), 225-239. Faysse, N., J.-D. Rinaudo, S. Bento, A. Richard-Ferroudji, M. Errahj, M. Varanda, A. Imache, M. Dionnet, D. Rollin, P. Garin, M. Kuper, L. Maton, and M. Montginoul, 2012: Participatory analysis for adaptation to climate change in Mediterranean agricultural systems: possible choices in process design. Regional Environmental Change,(Published Online 20 October 2013), 1-14. Fermont, A.M., P.J.A. van Asten, and K.E. Giller, 2008: Increasing land pressure in East Africa: The changing role of cassava and consequences for sustainability of farming systems. Agriculture, Ecosystems and Environment, 128(4), 239-250. Ferrari, M.C.O., D.L. Dixson, P.L. Munday, M.I. Mccormick, M.G. Meekan, A. Sih, and D.P. Chivers, 2011: Intrageneric variation in antipredator responses of coral reef fishes affected by ocean acidification: Implications for climate change projections on marine communities. Global Change Biology, 17(9), 2980-2986. Figueiredo, P. and P.E. Perkins, 2012: Women and water management in times of climate change: participatory and inclusive processes. Journal of Cleaner Production, In Press, Corrected Proof. Findlay, A.M., 2011: Migrant destinations in an era of environmental change. Global Environmental Change, 21(SUPPL. 1), S50-S58. Fisher, M., M. Chaudhury, and B. McCusker, 2010: Do forests help rural households adapt to climate variability? Evidence from Southern Malawi. World Development, 38(9), 1241-1250. Flam, K.H. and J.B. Skjaerseth, 2009: Does adequate financing exist for adaptation in developing countries? Climate Policy, 9(1), 109-114. Foden, W., G.F. Midgley, G. Hughes, W.J. Bond, W. Thuiller, M.T. Hoffman, P. Kaleme, L.G. Underhill, A. Rebelo, and L. Hannah, 2007: A changing climate is eroding the geographical range of the Namib Desert tree Aloe through population declines and dispersal lags. Diversity and Distributions, 13(5), 645-653. Folke, C., T. Hahn, P. Olsson, and J. Norberg, 2005: Adaptive governance of social-ecological systems. Annual Review of Environment and Resources, 30, 441-473. Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Folke, C., S. Carpenter, T. Elmqvist, L. Gunderson, C.S. Holling, B. Walker, J. Bengtsson, F. Berkes, J. Colding, K. Danell, M. Falkenmark, L. Gordon, R. Kasperson, N. Kautsky, A. Kinzig, S. Levin, K. Mäler, F. Moberg, L. Ohlsson, P. Olsson, E. Ostrom, W. Reid, J. Rockström, H. Savenije, and U. Svedin, 2002: Resilience and Sustainable Development: Building Adaptive Capacity in a World of Transformations. Scientific Background Paper on Resilience for the process of The World Summit on Sustainable Development. Environmental Advisory Council to the Swedish Government, Stockholm, Sweden, pp. 34. Fontaine, B., P. Roucou, and P.-A. Monerie, 2011: Changes in the African monsoon region at medium-term time horizon using 12 AR4 coupled models under the A1B emissions scenario. Atmospheric Science Letters, 12(1), 83-88. Ford, J.D., W. Vanderbilt, and L. Berrang-Ford, 2012: Authorship in IPCC AR5 and its implications for content: Climate change and Indigenous populations in WGII. Climatic Change, 113(2), 201-213. Foresight, 2011: Migration and global environmental change: future challenges and opportunities. Final Project Report. The Government Office for Science, London, UK, pp. 234. Fox, S. and K. Hoelscher, 2010: The political economy of social violence: theory and evidence from a cross-country study. Crisis States Working Papers Series No. 2, April 2010. Crisis States Research Centre, London School of Economics, London, UK, pp. 24. Frankenberger, T.R., T. Spangle, S. Nelson, and M. Langworthy, 2012: Enhancing Resilience to Food Insecurity amid Protracted Crisis. Proceedings of High Level Expert Forum on food insecurity in protracted crisis, 13-14 September 2012, Rome, Italy, pp. 16. Freier, K.P., R. Bruggemann, J. Scheffran, M. Finckh, and U.A. Schneider, 2012: Assessing the predictability of future livelihood strategies of pastoralists in semi-arid Morocco under climate change. Technological Forecasting and Social Change, 79(2), 371-382. Friedman, J. and N. Schady, 2009: How many more infants are likely to die in Africa as a result of the global financial crisis?. Policy Research Working Paper No. 5023. Development Reserach Group, The World Bank, Washington, D.C., pp. 16. Fröde, A., M. Scholze, and N. Manasfi, 2013: Taking a climate perspective on development: GIZ's climate proofing for development approach. Climate and Development, 5(2), 160-164. Funk, C., G. Eilerts, J. Verdin, J. Rowland, and M. Marshall, 2011: A climate trend analysis of Sudan. U.S. Geological Survey Fact Sheet 2011-3072, pp. 6. Funk, C., J. Michaelsen, and M. Marshall, 2012: Mapping recent decadal climate variations in precipitation and temperature across Eastern Africa and the Sahel. In: Remote Sensing of Drought: Innovative Monitoring Approaches. [Wardlow, B.D., M.C. Anderson, and J.P. Verdin (eds.)]. CRC Press, Boca Raton, FL, USA, pp. 331-358. Funk, C., M.D. Dettinger, J.C. Michaelsen, J.P. Verdin, M.E. Brown, M. Barlow, and A. Hoell, 2008: Warming of the Indian Ocean threatens eastern and southern African food security but could be mitigated by agricultural development. Proceedings of the National Academy of Sciences of the United States of America, 105(32), 11081-11086. Galloway McLean, K., 2010: Advance Guard: Climate Change Impacts, Adaptation, Mitigation and Indigenous Peoples A Compendium of Case Studies. United Nations University, Institute of Advanced Studies - Traditional Knowledge Initiative, Darwin, Australia, pp. 124. Galvin, K.A., 2009: Transitions: Pastoralists living with change. Annual Review of Anthropology, 38, 185-198. Gao, X. and F. Giorgi, 2008: Increased aridity in the Mediterranean region under greenhouse gas forcing estimated from high resolution simulations with a regional climate model. Global and Planetary Change, 62(3-4), 195- 209. García-Ruiz, J.M., J.I. López-Moreno, S.M. Vicente-Serrano, T. Lasanta-Martínez, and S. Beguería, 2011: Mediterranean water resources in a global change scenario. Earth-Science Reviews, 105(3-4), 121-139. Garrity, D.P., F.K. Akinnifesi, O.C. Ajayi, S.G. Weldesemayat, J.G. Mowo, A. Kalinganire, M. Larwanou, and J. Bayala, 2010: Evergreen Agriculture: A robust approach to sustainable food security in Africa. Food Security, 2(3), 197-214. Gbetibouo, G.A., 2009: Understanding farmers' perceptions and adaptations to climate change and variability: The case of the Limpopo Basin, South Africa. IFPRI Discussion Paper 00849. International Food Policy Research Institute (IFPRI) Environment and Production Technology Division, Washington, DC, pp. 43. Gbetibouo, G.A., R.M. Hassan, and C. Ringler, 2010a: Modelling farmers' adaptation strategies for climate change and variability: The case of the Limpopo Basin, South Africa. Agrekon, 49(2), 217-234. Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Gbetibouo, G.A., C. Ringler, and R. Hassan, 2010b: Vulnerability of the South African farming sector to climate change and variability: An indicator approach. Natural Resources Forum, 34(3), 175-187. GEF Evaluation Office, 2011: Evaluation of the Special Climate Change Fund (SCCF). Unedited Version of Final Report. Global Environmental Facility (GEF), pp. 62. Gemenne, F., 2011a: Climate-induced population displacements in a 4°C+ world. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1934), 182-195. Gemenne, F., 2011b: Why the numbers don't add up: A review of estimates and predictions of people displaced by environmental changes. Global Environmental Change, 21(SUPPL. 1), S41-S49. German Commission for UNESCO, 2011: For life, for the future. Biosphere reserves and climate change. A collection of good practice case studies. [Möller, L. (ed.)]. German Commission for UNESCO (DUK), Bonn, Germany, pp. 80. German, L., G.C. Schoneveld, and D. Gumbo, 2011: The local social and environmental impacts of smallholder- based biofuel investments in Zambia. Ecology and Society, 16(4), 12. Gething, P.W., D.L. Smith, A.P. Patil, A.J. Tatem, R.W. Snow, and S.I. Hay, 2010: Climate change and the global malaria recession. Nature, 465(7296), 342-345. Gijsbers, H.J.M., J.J. Kessler, and M.K. Knevel, 1994: Dynamics and natural regeneration of woody species in farmed parklands in the Sahel region (Province of Passore, Burkina Faso). Forest Ecology and Management, 64(1), 1-12. Gilioli, G. and L. Mariani, 2011: Sensitivity of Anopheles gambiae population dynamics to meteo-hydrological variability: A mechanistic approach. Malaria Journal, 10, 294. Giorgi, F. and P. Lionello, 2008: Climate change projections for the Mediterranean region. Global and Planetary Change, 63(2-3), 90-104. Giorgi, F., C. Jones, and G.R. Asrar, 2009: Addressing climate information needs at the regional level: the CORDEX framework. WMO Bulletin, 58(3), 175-183. Girot, P., C. Ehrhart, and J. Oglethorpe, 2012: Integrating Community and Ecosystem-Based Approaches in Climate Change Adaptation Responses. Briefing Paper. Ecosystem & Livelihoods Adaptation Networks (ELAN), pp. 19. Githeko, A.K., E.N. Ototo, and Y. Guiyun, 2012: Progress towards understanding the ecology and epidemiology of malaria in the western Kenya highlands: Opportunities and challenges for control under climate change risk. Acta Tropica, 121(1), 19-25. Gleditsch, N.P., 2012: Whither the weather? Climate change and conflict. Journal of Peace Research, 49(1), 3-9. Godfray, H.C.J., J.R. Beddington, I.R. Crute, L. Haddad, D. Lawrence, J.F. Muir, J. Pretty, S. Robinson, S.M. Thomas, and C. Toulmin, 2010: Food security: The challenge of feeding 9 billion people. Science, 327(5967), 812-818. Godfrey, A., E. Le Roux-Rutledge, S. Cooke, and M. Burton, 2010: Africa Talks Climate: The public understanding of climate change in ten countries. Executive Summary. BBC World Service Trust and British Council, London, UK, pp. 11. Goh, A.H.X., 2012: A literature review of the gender-differentiated impacts of climate change on women s and men s assets and well-being in developing countries. CAPRi Working Paper No. 106. International Food Policy Research Institute (IFPRI), Washington, DC, USA, pp. 38. Goklany, I.M., 2007: Integrated strategies to reduce vulnerability and advance adaptation, mitigation, and sustainable development. Mitigation and Adaptation Strategies for Global Change, 12(5), 755-786. Goldstone, J.A., 2002: Population and security: how demographic change can lead to violent conflict. Journal of International Affairs, 56(1), 3-22. Gomez-Echeverri, L., 2010: National Funding Entities: Their role in the transition to a new paradigm of global cooperation on climate change. ECBI Policy Report. European Capacity Building Initiative (ECBI), Oxford, UK, pp. 40. Gonzalez, P., R.P. Neilson, J.M. Lenihan, and R.J. Drapek, 2010: Global patterns in the vulnerability of ecosystems to vegetation shifts due to climate change. Global Ecology and Biogeography, 19(6), 755-768. Gonzalez, P., C.J. Tucker, and H. Sy, 2012: Tree density and species decline in the African Sahel attributable to climate. Journal of Arid Environments, 78, 55-64. Goulden, M., D. Conway, and A. Persechino, 2009a: Adaptation to climate change in international river basins in Africa: A review. Hydrological Sciences Journal, 54(5), 805-828. Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Goulden, M., L.O. Naess, K. Vincent, and W.N. Adger, 2009b: Accessing Diversification, Networks and Traditional Resource Management as Adaptations to Climate Extremes. In: Adapting to Climate Change: Thresholds, Values, Governance. [Adger, W.N., I. Lorenzoni, and K. O Brien (eds.)]. Cambridge University Press, Cambridge, UK, pp. 448-464. Government of Uganda and WFP, 2010: Project Document on Partnership between the Office of the Prime Minister of Uganda and the United Nations World Food Programme in the Implementation of the Livelihoods Program under the Framework of the Second Northern Uganda Social Action Fund (NUSAF2) in Karamoja Sub-Region. pp. 25. Grab, S. and A. Craparo, 2011: Advance of apple and pear tree full bloom dates in response to climate change in the southwestern Cape, South Africa: 1973-2009. Agricultural and Forest Meteorology, 151(3), 406-413. Grace, K., F. Davenport, C. Funk, and A.M. Lerner, 2012: Child malnutrition and climate in Sub-Saharan Africa: An analysis of recent trends in Kenya. Applied Geography, 35(1), 405-413. Gray, C.L., 2011: Soil quality and human migration in Kenya and Uganda. Global Environmental Change, 21(2), 421-430. Greene, A.M., A. Giannini, and S.E. Zebiak, 2009: Drought return times in the Sahel: A question of attribution. Geophysical Research Letters, 36(12). Grothmann, T. and A. Patt, 2005: Adaptive capacity and human cognition: The process of individual adaptation to climate change. Global Environmental Change, 15(3), 199-213. Guariguata, M.R., J.P. Cornelius, B. Locatelli, C. Forner, and G.A. Sánchez-Azofeifa, 2008: Mitigation needs adaptation: Tropical forestry and climate change. Mitigation and Adaptation Strategies for Global Change, 13(8), 793-808. Guthiga, P. and A. Newsham, 2011: Meteorologists Meeting Rainmakers: Indigenous Knowledge and Climate Policy Processes in Kenya. IDS Bulletin, 42(3), 104-109. Haan, N., S. Devereux, and D. Maxwell, 2012: Global implications of Somalia 2011 for famine prevention, mitigation and response. Global Food Security, 1, 74-79. Haarsma, R.J., F.M. Selten, S.L. Weber, and M. Kliphuis, 2005: Sahel rainfall variability and response to greenhouse warming. Geophysical Research Letters, 32(17), 1-4. Hadley, C., D.A. Linzer, T. Belachew, A.G. Mariam, F. Tessema, and D. Lindstrom, 2011: Household capacities, vulnerabilities and food insecurity: Shifts in food insecurity in urban and rural Ethiopia during the 2008 food crisis. Social Science and Medicine, 73(10), 1534-1542. Haensler, A., F. Saeed, and D. Jacob, 2013: Assessing the robustness of projected precipitation changes over central Africa on the basis of a multitude of global and regional climate projections. Climatic Change, 1-15. Haglund, E., J. Ndjeunga, L. Snook, and D. Pasternak, 2011: Dry land tree management for improved household livelihoods: Farmer managed natural regeneration in Niger. Journal of Environmental Management, 92(7), 1696-1705. Hahn, M.B., A.M. Riederer, and S.O. Foster, 2009: The Livelihood Vulnerability Index: A pragmatic approach to assessing risks from climate variability and change-A case study in Mozambique. Global Environmental Change, 19(1), 74-88. Haites, E., 2011: Climate change finance. Climate Policy, 11(3), 963-969. Halsnaes, K. and S. Trarup, 2009: Development and climate change: A mainstreaming approach for assessing economic, social, and environmental impacts of adaptation measures. Environmental Management, 43(5), 765- 778. Hambira, W.L., 2011: Screening for climate change vulnerability in Botswana's tourism sector in a bid to explore suitable adaptation measures and policy implications: A case study of the Okavango Delta. International Journal of Tourism Policy, 4(1), 51-65. Hamisi, H.I., M. Tumbo, E. Kalumanga, and P. Yanda, 2012: Crisis in the wetlands: Combined stresses in a changing climate - Experience from Tanzania. Climate and Development, 4(1), 5-15. Hanff, E., M.-H. Dabat, and J. Blin, 2011: Are biofuels an efficient technology for generating sustainable development in oil-dependent African nations? A macroeconomic assessment of the opportunities and impacts in Burkina Faso. Renewable and Sustainable Energy Reviews, 15(5), 2199-2209. Harmeling, S., S. Kreft, and S.C. Rai, 2011: Institutions for adaptation: towards an effective multi-level interplay. Germanwatch e.V. and WWF International, Bonn, Germany and Gland, Switzerland, pp. 44. Hartmann, I. and A.J. Sugulle, 2009: The impact of climate change on pastoral societies of Somaliland. Candlelight for Health, Education & Environment, pp. 62. Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Harvey, B., J. Ensor, L. Carlile, B. Garside, Z. Patterson, and L.O. Naess, 2012: Climate change communication and social learning - Review and strategy development for CCAFS. CCAFS Working Paper No. 22. CGIAR Research Program on Climate Change, Agriculture and Food Security (CCAFS), Copenhagen, Denmark, pp. 51. Harvey, B., 2011: Climate airwaves: Community radio, action research, and advocacy for climate justice in Ghana. International Journal of Communication, 5(1), 2035-2058. Hashizume, M., T. Terao, and N. Minakawa, 2009: The Indian Ocean Dipole and malaria risk in the highlands of western Kenya. Proceedings of the National Academy of Sciences of the United States of America, 106(6), 1857-1862. Hashizume, M., L.F. Chaves, and N. Minakawa, 2012: Indian Ocean Dipole drives malaria resurgence in East African highlands. Scientific Reports, 2, 269. Hassaan, M.A. and M.A. Abdrabo, 2013: Vulnerability of the Nile Delta coastal areas to inundation by sea level rise. Environmental Monitoring and Assessment, 185(8), 6607-6616. Hecky, R.E., R. Mugidde, P.S. Ramlal, M.R. Talbot, and G.W. Kling, 2010: Multiple stressors cause rapid ecosystem change in Lake Victoria. Freshwater Biology, 55(SUPPL. 1), 19-42. Hegazy, A.K., M.A. Medany, H.F. Kabiel, and M.M. Maez, 2008: Spatial and temporal projected distribution of four crop plants in Egypt. Natural Resources Forum, 32(4), 316-326. Heidecke, C. and T. Heckelei, 2010: Impacts of changing water inflow distributions on irrigation and farm income along the Drâa River in Morocco. Agricultural Economics, 41(2), 135-149. Hein, L., M.J. Metzger, and R. Leemans, 2009: The local impacts of climate change in the Ferlo, Western Sahel. Climatic Change, 93(3-4), 465-483. Hellmuth, M.E., A. Moorhead, M.C. Thomson, and J. Williams (eds.), 2007: Climate Risk Management in Africa: Learning from Practice International Research Institute for Climate and Society (IRI), Columbia University, New York, USA, pp. 104. Hellmuth, M.E., D.E. Osgood, U. Hess, A. Moorhead, and H. Bhojwani (eds.), 2009: Index insurance and climate risk: prospects for development and disaster management. In: Climate and Society No. 2. International Research Institute for Climate and Society (IRI), Columbia University, New York, USA, pp. 112. Heltberg, R., P.B. Siegel, and S.L. Jorgensen, 2009: Addressing human vulnerability to climate change: Toward a 'no-regrets' approach. Global Environmental Change, 19(1), 89-99. Hemp, A., 2009: Climate change and its impact on the forests of Kilimanjaro. African Journal of Ecology, 47(SUPPL. 1), 3-10. Hendrix, C.S. and I. Salehyan, 2012: Climate change, rainfall, and social conflict in Africa. Journal of Peace Research, 49(1), 35-50. Hertel, T.W., M.B. Burke, and D.B. Lobell, 2010: The poverty implications of climate-induced crop yield changes by 2030. Global Environmental Change, 20(4), 577-585. Hession, S.L. and N. Moore, 2011: A spatial regression analysis of the influence of topography on monthly rainfall in East Africa. International Journal of Climatology, 31(10), 1440-1456. Hewitson, B.C. and R.G. Crane, 2006: Consensus between GCM climate change projections with empirical downscaling: Precipitation downscaling over South Africa. International Journal of Climatology, 26(10), 1315- 1337. Hewitson, B., C. Lennard, G. Nikulin, and C. Jones, 2012: CORDEX-Africa: a unique opportunity for science and capacity building. CLIVAR Exchanges, 17(3), 6-7. Hightower, A., C. Kinkade, P.M. Nguku, A. Anyangu, D. Mutonga, J. Omolo, M.K. Njenga, D.R. Feikin, D. Schnabel, M. Ombok, and R.F. Breiman, 2012: Relationship of climate, geography, and geology to the incidence of rift valley fever in Kenya during the 2006-2007 outbreak. American Journal of Tropical Medicine and Hygiene, 86(2), 373-380. Himeidan, Y.E. and E.J. Kweka, 2012: Malaria in East African highlands during the past 30 years: Impact of environmental changes. Frontiers in Physiology, 3, 315. Hirsch Hadorn, G., S. Biber-Klemm, W. Grossenbacher-Mansuy, H. Hoffmann-Riem, D. Joye, C. Pohl, U. Wiesmann, and E. Zemp, 2008: The Emergence of Transdisciplinarity as a Form of Research. In: Handbook of Transdisciplinary Research. [Hirsch Hadorn, G., H. Hoffmann-Riem, S. Biber-Klemm, W. Grossenbacher- Mansuy, D. Joye, C. Pohl et al. (eds.)]. Springer Netherlands, Dordrecht, pp. 19-39. Hisali, E., P. Birungi, and F. Buyinza, 2011: Adaptation to climate change in Uganda: Evidence from micro level data. Global Environmental Change, 21(4), 1245-1261. Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hitchcock, R.K., 2009: From local to global: perceptions and realities of environmental change among Kalahari San. In: Anthropology and climate change: from encounters to actions. [Crate, S.A. and M. Nuttall (eds.)]. Left Coast Press, Walnut Creek, CA, USA, pp. 250-264. Hockey, P.A.R. and G.F. Midgley, 2009: Avian range changes and climate change: A cautionary tale from the Cape Peninsula. Ostrich, 80(1), 29-34. Hockey, P.A.R., C. Sirami, A.R. Ridley, G.F. Midgley, and H.A. Babiker, 2011: Interrogating recent range changes in South African birds: Confounding signals from land use and climate change present a challenge for attribution. Diversity and Distributions, 17(2), 254-261. Hodgson, A., T. Smith, S. Gagneux, M. Adjuik, G. Pluschke, N.K. Mensah, F. Binka, and B. Genton, 2001: Risk factors for meningococcal meningitis in northern Ghana. Transactions of the Royal Society of Tropical Medicine and Hygiene, 95(5), 477-480. Hoegh-Guldberg, O., P.J. Mumby, A.J. Hooten, R.S. Steneck, P. Greenfield, E. Gomez, C.D. Harvell, P.F. Sale, A.J. Edwards, K. Caldeira, N. Knowlton, C.M. Eakin, R. Iglesias-Prieto, N. Muthiga, R.H. Bradbury, A. Dubi, and M.E. Hatziolos, 2007: Coral reefs under rapid climate change and ocean acidification. Science, 318(5857), 1737-1742. Hoerling, M., J. Hurrell, J. Eischeid, and A. Phillips, 2006: Detection and attribution of twentieth-century northern and southern African rainfall change. Journal of Climate, 19(16), 3989-4008. Hoffman, M.T., M.D. Cramer, L. Gillson, and M. Wallace, 2011: Pan evaporation and wind run decline in the Cape Floristic Region of South Africa (1974-2005): implications for vegetation responses to climate change. Climatic Change, 109(3-4), 437-452. Hsiang, S.M., M. Burke, and E. Miguel, 2013: Quantifying the influence of climate on human conflict. Science, 341(6151). Huang, S.-Y., Z.-H. Deng, Q.-M. Zhang, R.-X. Lin, X.-C. Zhang, L.-C. Huo, J.-L. Wang, and C.-W. Ruan, 2011: Endemic situation of schistosomiasis in Guangdong Province from 2004 to 2009. Chinese Journal of Schistosomiasis Control, 23(2), 197-198+201. Huang, Y. and L. Manderson, 1992: Schistosomiasis and the social patterning of infection. Acta Tropica, 51(3-4), 175-194. Hulme, M., R. Doherty, T. Ngara, M. New, and D. Lister, 2001: African climate change: 1900-2100. Climate Research, 17(2), 145-168. Hunde, M., 2012: Integrating community based disaster risk reduction and climate change adaptation into local development planning. Proceedings of Sixth International Conference on Community-Based Adaptation, 16-22, April 2012, Hanoi, Vietnam, pp. 20-21. Hunter, J.M., L. Rey, K.Y. Chu, E.O. Adekolu-John, and K.E. Mott, 1993: Parasitic diseases in water resources development: The need for intersectoral negotiation. World Health Organization (WHO), Geneva, Switzerland, pp. 152. Huq, S., 2011: Improving information for community-based adaptation. IIED Opinion: Lessons from adaptation in practice, October 2011. International Institute for Environment and Development (IIED), London, UK, pp. 2. IDMC, 2011: Displacement due to natural hazard-induced disaster: Global estimates for 2009 and 2010. Internal Displacement Monitoring Centre (IDMC), Geneva, Switzerland, pp. 30. IDS, 2012: Innovative approaches to gender and food security: changing attitudes, changing behaviours. Food Security Insights Issue 82, January 2012. IDS Knowledge Services, Brighton, UK, pp. 8. IFAD, 2010: Rural poverty report 2011. New realities, new challenges: new opportunities for tomorrow's generations. International Fund for Agricultural Development (IFAD), Rome, Italy, pp. 317. IFAD, WFP, and FAO, 2012: The State of Food Insecurity in the World 2012. Economic growth is necessary but not sufficient to accelerate reduction of hunger and malnutrition. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy, pp. 61. Ifejika Speranza, C., 2010: Drought coping and adaptation strategies: Understanding adaptations to climate change in agro-pastoral livestock production in Makueni district, Kenya. European Journal of Development Research, 22(5), 623-642. Ifejika Speranza, C., B. Kiteme, P. Ambenje, U. Wiesmann, and S. Makali, 2010: Indigenous knowledge related to climate variability and change: Insights from droughts in semi-arid areas of former Makueni District, Kenya. Climatic Change, 100(2), 295-315. IPCC, 2007: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Solomon, S., Qin, D., Manning, M., Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chen, Z., Marquis, M., Averyt, K.B., Tignor, M. and Miller, H.L. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 996. IPCC, 2012: Summary for Policymakers. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX). Special Report of the Intergovernmental panel on Climate Change. [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al. (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 1-19. Itiveh, K.O. and G.R. Bigg, 2008: The variation of discharge entering the Niger Delta system, 1951-2000, and estimates of change under global warming. International Journal of Climatology, 28(5), 659-666. Jackson, M.C., L. Johansen, C. Furlong, A. Colson, and K.F. Sellers, 2010: Modelling the effect of climate change on prevalence of malaria in western Africa. Statistica Neerlandica, 64(4), 388-400. Jäger, J., J. Frühmann, S. Grünberger, and A. Vag, 2009: EACH-FOR. Environmental Change and Forced Migration Scenarios: D.3.4. Synthesis Report. pp. 81. Jalloh, A., B. Sarr, J. Kuiseu, H. Roy-Macauley, and P. Sereme, 2011a: Review of climate in West and Central Africa to inform farming systems research and development in the sub-humid and semi-arid agroecologies of the region. Conseil Ouest et Centre Africain pour la Recherche et le Development Agricoles/West and Central African Council for Agricultural Research and Development (CORAF/WECARD), Dakar, Senegal, pp. 53. Jalloh, A., H. Roy-Macauley, and J. Kuiseu, 2011b: Climate change research and development orientation framework: A CORAF/WECARD strategy for guiding climate change research and development in West and Central Africa. Conseil Ouest et Centre Africain pour la Recherche et le Development Agricoles/West and Central African Council for Agricultural Research and Development (CORAF/WECARD), Dakar, Senegal, pp. 46. James, R. and R. Washington, 2013: Changes in African temperature and precipitation associated with degrees of global warming. Climatic Change, 117(4), 859-872. Jankowska, M.M., D. Lopez-Carr, C. Funk, G.J. Husak, and Z.A. Chafe, 2012: Climate change and human health: Spatial modeling of water availability, malnutrition, and livelihoods in Mali, Africa. Applied Geography, 33(1), 4-15. Jaramillo, J., E. Muchugu, F.E. Vega, A. Davis, C. Borgemeister, and A. Chabi-Olaye, 2011: Some like it hot: The influence and implications of climate change on coffee berry borer (Hypothenemus hampei) and coffee production in East Africa. PLoS ONE, 6(9). Jarvis, A., J. Ramirez-Villegas, B.V.H. Campo, and C. Navarro-Racines, 2012: Is Cassava the Answer to African Climate Change Adaptation? Tropical Plant Biology, 5(1), 9-29. Jayne, T.S., B. Zulu, and J.J. Nijhoff, 2006: Stabilizing food markets in eastern and southern Africa. Food Policy, 31(4), 328-341. Jennings, S. and J. Magrath, 2009: What happened to the seasons?. Oxfam GB Research Report. Oxfam GB, pp. 26. Jima, D., M. Wondabeku, A. Alemu, A. Teferra, N. Awel, W. Deressa, A. Adissie, Z. Tadesse, T. Gebre, A.W. Mosher, F.O. Richards, and P.M. Graves, 2012: Analysis of malaria surveillance data in Ethiopia: What can be learned from the Integrated Disease Surveillance and Response System? Malaria Journal, 11, 330. Jobbins, G., 2011: Managing uncertainty in adaptation. CCAA Perspectives series. International Development Research Centre (IDRC) and Department for International Development (DFID), Ottawa, Canada, pp. 10. Jobin, W., 1999: Dams and disease: ecological design and health impacts of large dams, canals and irrigation systems. Taylor & Francis, London, UK and New York, NY, USA, pp. 600. Jones, L., E. Ludi, and S. Levine, 2010: Towards a characterisation of adaptive capacity: a framework for analysing adaptive capacity at the local level. ODI Background Note, December 2010. Overseas Development Institute (ODI), London, UK, pp. 8. Jones, L., 2012: Social barriers to adaptation: exploring implications and identifying options for adaptation policy across the SADC Region. In: Overcoming barriers to climate change adaptation implementation in Southern Africa. [Masters, L. and L. Duff (eds.)]. Africa Institute of South Africa, Pretoria, South Africa, pp. 41-60. Jones, C., F. Giorgi, and G. Asrar, 2011: The Coordinated Regional Downscaling Experiment: CORDEX - An international downscaling link to CMIP5. CLIVAR Exchanges, 16(2), 34-40. Jones, H.P., D.G. Hole, and E.S. Zavaleta, 2012: Harnessing nature to help people adapt to climate change. Nature Climate Change, 2(7), 504-509. Jones, L. and E. Boyd, 2011: Exploring social barriers to adaptation: Insights from Western Nepal. Global Environmental Change, 21(4), 1262-1274. Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Jones, P.D. and A. Moberg, 2003: Hemispheric and large-scale surface air temperature variations: An extensive revision and an update to 2001. Journal of Climate, 16(2), 206-223. Jones, P.G. and P.K. Thornton, 2009: Croppers to livestock keepers: livelihood transitions to 2050 in Africa due to climate change. Environmental Science and Policy, 12(4), 427-437. Joshi, M., E. Hawkins, R. Sutton, J. Lowe, and D. Frame, 2011: Projections of when temperature change will exceed 2°C above pre-industrial levels. Nature Climate Change, 1(8), 407-412. Kabubo-Mariara, J., 2008: Climate change adaptation and livestock activity choices in Kenya: An economic analysis. Natural Resources Forum, 32(2), 131-141. Kalame, F.B., M. Brockhaus, J. Nkem, M. Idinoba, and M. Kanninen, 2008: Using forest ecosystem goods and services for climate change adaptation in Burkina Faso and Ghana: Policy gaps and constraints. Proceedings of International conferences on adaptation of forests and forest management to changing climate with emphasis on forest health: A review of science, policies and practices. Book of abstracts and preliminary programme, August 25-28 2008, Umea, Sweden, pp. 109. Kalame, F.B., O. Luukkanen, and M. Kanninen, 2011: Making the national adaptation programme of action (NAPA) more responsive to the livelihood needs of tree planting farmers, drawing on previous experience in Dryland Sudan. Forests, 2(4), 948-960. Kälin, W. and N. Schrepfer, 2012: Protecting people crossing borders in the context of climate change: normative gaps and possible approaches. Legal and Protection Policy Research Series. UN High Commissioner for Refugees, pp. 80. Kalognomou, E., C. Lennard, M. Shongwe, I. Pinto, A. Favre, M. Kent, B. Hewitson, A. Dosio, G. Nikulin, H. Panitz, and M. Büchner, 2013: A diagnostic evaluation of precipitation in CORDEX models over southern Africa. Journal of Climate, Early Online Release. Kandji, S.T., L. Verchot, and J. Mackensen, 2006: Climate change and variability in the Sahel region: impacts and adaptation strategies in the agricultural sector. United Nations Environment Programme (UNEP) and World Agroforestry Centre, Nairobi, Kenya, pp. 48. Karlberg, L., J. Rockström, J.G. Annandale, and J.M. Steyn, 2007: Low-cost drip irrigation-A suitable technology for southern Africa?. An example with tomatoes using saline irrigation water. Agricultural Water Management, 89(1-2), 59-70. Kassam, A., T. Friedrich, R. Derpsch, R. Lahmar, R. Mrabet, G. Basch, E.J. González-Sánchez, and R. Serraj, 2012: Conservation agriculture in the dry Mediterranean climate. Field Crops Research, 132, 7-17. Kay, G. and R. Washington, 2008: Future southern African summer rainfall variability related to a southwest Indian Ocean dipole in HadCM3. Geophysical Research Letters, 35(12). Kebede, A.S., S. Brown, and R.J. Nicholls, 2010: Synthesis Report: The Implications of Climate Change and Sea- Level Rise in Tanzania - The Coastal Zones. Report submitted to the Stockholm Environment Institute, 5 November 2010, pp. 32. Kebede, A.S. and R.J. Nicholls, 2011: Population and asset exposure to coastal flooding in Dar es Salaam (Tanzania): vulnerability to climate extremes. Report submitted to Global Climate Adaptation Partnership (GCAP), pp. 26. Kebede, A.S. and R.J. Nicholls, 2012: Exposure and vulnerability to climate extremes: Population and asset exposure to coastal flooding in Dar es Salaam, Tanzania. Regional Environmental Change, 12(1), 81-94. Kebede, A.S., R.J. Nicholls, S. Hanson, and M. Mokrech, 2012: Impacts of climate change and sea-level rise: A preliminary case study of Mombasa, Kenya. Journal of Coastal Research, 28(1 A), 8-19. Kelly-Hope, L.A., J. Hemingway, and F.E. McKenzie, 2009: Environmental factors associated with the malaria vectors Anopheles gambiae and Anopheles funestus in Kenya. Malaria Journal, 8, 268. Kemp-Benedict, E. and W.K. Agyemang-Bonsu, 2008: The Akropong approach to multi-sector project planning. Futures, 40(9), 834-840. Kerrou, J., P. Renard, and J. Tarhouni, 2010: Status of the Korba groundwater resources (Tunisia): Observations and three-dimensional modelling of seawater intrusion. Hydrogeology Journal, 18(5), 1173-1190. Kienberger, S., 2012: Spatial modelling of social and economic vulnerability to floods at the district level in Búzi, Mozambique. Natural Hazards, 64(3), 2001-2019. Kim, J., D.E. Waliser, C.A. Mattmann, C.E. Goodale, A.F. Hart, P.A. Zimdars, D.J. Crichton, C. Jones, G. Nikulin, B. Hewitson, C. Jack, C. Lennard, and A. Favre, 2013: Evaluation of the CORDEX-Africa multi-RCM hindcast: systematic model errors. Climate Dynamics, 1-14. Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kimutai, A., P. Kamau Ngure, W. Kiprotich Tonui, M. Muita Gicheru, and L. Bonareri Nyamwamu, 2006: Leishmaniasis in Northern and Western Africa: A review. African Journal of Infectious Diseases, 3(1), 14-25. Kingsford, R.T., H.C. Biggs, and S.R. Pollard, 2011: Strategic Adaptive Management in freshwater protected areas and their rivers. Biological Conservation, 144(4), 1194-1203. Kingston, D.G. and R.G. Taylor, 2010: Sources of uncertainty in climate change impacts on river discharge and groundwater in a headwater catchment of the Upper Nile Basin, Uganda. Hydrology and Earth System Sciences, 14(7), 1297-1308. Kithiia, J. and R. Dowling, 2010: An integrated city-level planning process to address the impacts of climate change in Kenya: The case of Mombasa. Cities, 27(6), 466-475. Kithiia, J., 2011: Climate change risk responses in East African cities: Need, barriers and opportunities. Current Opinion in Environmental Sustainability, 3(3), 176-180. Kithiia, J. and A. Lyth, 2011: Urban wildscapes and green spaces in Mombasa and their potential contribution to climate change adaptation and mitigation. Environment and Urbanization, 23(1), 251-265. Klempa, B., 2009: Hantaviruses and climate change. Clinical Microbiology and Infection, 15(6), 518-523. Kniveton, D.R., R. Layberry, C.J.R. Williams, and M. Peck, 2009: Trends in the start of the wet season over Africa. International Journal of Climatology, 29(9), 1216-1225. Koch, I.C., C. Vogel, and Z. Patel, 2007: Institutional dynamics and climate change adaptation in South Africa. Mitigation and Adaptation Strategies for Global Change, 12(8), 1323-1339. Koelle, B. and W. Annecke, 2011: Community based climate change adaptation (CBA). Lessons learnt from the Community Based Adaptation in Africa (CBAA). Adaptation and beyond No. 1. Indigo development & change, Nieuwoudtville, South Africa, pp. 4. Koelle, K., X. Rodó, M. Pascual, M. Yunus, and G. Mostafa, 2005: Refractory periods and climate forcing in cholera dynamics. Nature, 436(7051), 696-700. Kolmannskog, V., 2010: Climate change, human mobility, and protection: Initial evidence from Africa. Refugee Survey Quarterly, 29(3), 103-119. Koné, B., M. Feagan, Y.A. Houenou, N. Brou, P.V. Houenou, B. Fayomi, E. Ngnikam, G. Cissé, J. Spiegel, and E. Kouassi, 2011: Facilitating the Relationship Between Researchers and Policy-Makers: Experiences from Three Ecohealth Projects in West and Central Africa. EcoHealth, 8(4), 413-417. Krishnamurty, J., 2011: Employment policies and disaster risk reduction. Global Assessment Report on Disaster Risk Reduction. UNISDR and ILO, Geneva, Switzerland, pp. 27. Kruger, A.C. and S. Shongwe, 2004: Temperature trends in South Africa: 1960-2003. International Journal of Climatology, 24(15), 1929-1945. Kruger, A.C. and S.S. Sekele, 2012: Trends in extreme temperature indices in South Africa: 1962-2009. International Journal of Climatology, 33(3), 661-676. Kuffner, I.B., A.J. Andersson, P.L. Jokiel, K.S. Rodgers, and F.T. MacKenzie, 2008: Decreased abundance of crustose coralline algae due to ocean acidification. Nature Geoscience, 1(2), 114-117. Kulkarni, M.A., R.E. Desrochers, and J.T. Kerr, 2010: High resolution niche models of malaria vectors in Northern Tanzania: A new capacity to predict malaria risk? PLoS ONE, 5(2). Kumar, N. and A.R. Quisumbing, 2011: Gendered Impacts of the 2007 08 Food Price Crisis: Evidence Using Panel Data from Rural Ethiopia. IFPRI Discussion Paper 01093. International Food Policy Research Institute (IFPRI), Poverty, Health, and Nutrition Division, Washington, D.C., USA, pp. 24. Kumssa, A. and J.F. Jones, 2010: Climate change and human security in Africa. International Journal of Sustainable Development and World Ecology, 17(6), 453-461. Kundzewicz, Z.W., 2011: Comparative assessment: fact or fiction?. Paper to be presented at the Workshop Including longterm climate change in hydrologic design", November 21, 2011, Washington, D.C., USA, pp. 17. Kunstmann, H., G. Jung, S. Wagner, and H. Clottey, 2008: Integration of atmospheric sciences and hydrology for the development of decision support systems in sustainable water management. Physics and Chemistry of the Earth, 33(1-2), 165-174. Kutsch, W.L., L. Merbold, W. Ziegler, M.M. Mukelabai, M. Muchinda, O. Kolle, and R.J. Scholes, 2011: The charcoal trap: Miombo forests and the energy needs of people. Carbon Balance and Management, 6. Läderach, P., A. Eitzinger, O. Ovalle, J. Ramirez, and A. Jarvis, 2010: Climate change adaptation and mitigation in the Kenyan coffee sector. Final Report. International Center for Tropical Agriculture (CIAT), Cali, Colombia, pp. 42. Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Läderach, P., A. Eitzinger, A. Martinez, and N. Castro, 2011a: Predicting the impact of climate change on cashew growing regions in Ghana and Cote d Ivoire. Climate Change, Agriculture and Food Security Programme (CCAFS) and International Center for Tropical Agriculture (CIAT), Managua, Nicaragua, pp. 29. Läderach, P., A. Eitzinger, A. Martinez, and N. Castro, 2011b: Predicting the impact of climate change on cotton growing regions in Ghana and Cote d Ivoire. Final report. Climate Change, Agriculture and Food Security Programme (CCAFS) and International Center for Tropical Agriculture (CIAT), Cali, Colombia, pp. 34. Läderach, P., A. Eitzinger, A. Martinez, and N. Castro, 2011c: Predicting the impact of climate change on the cocoa-growing regions in Ghana and Cote d Ivoire. Climate Change, Agriculture and Food Security Programme (CCAFS) and International Center for Tropical Agriculture (CIAT), Managua, Nicaragua, pp. 29. Lafontaine, A., Adejuwon J.O., P.N. Dearden, and G. Quesne, 2012: Final Evaluation of the IDRC/DFID Climate Change Adaptation in Africa Programme. Final Report, pp. 111. Lam, V.W.Y., W.W.L. Cheung, W. Swartz, and U.R. Sumaila, 2012: Climate Change Impacts on Fisheries in West Africa: Implications for Economic, Food and Nutritional Security. African Journal of Marine Science, 34(1), 103-117. Laprise, R., L. Hernández-Díaz, K. Tete, L. Sushama, L. ©eparovic, A. Martynov, K. Winger, and M. Valin, 2013: Climate projections over CORDEX Africa domain using the fifth-generation Canadian Regional Climate Model (CRCM5). Climate Dynamics,(Published online 22-01-2013), 1-28. Larwanou, M. and M. Saadou, 2011: The role of human interventions in tree dynamics and environmental rehabilitation in the Sahel zone of Niger. Journal of Arid Environments, 75(2), 194-200. Laube, W., B. Schraven, and M. Awo, 2012: Smallholder adaptation to climate change: Dynamics and limits in Northern Ghana. Climatic Change, 111(3), 753-774. LDC Expert Group, 2011: Best practices and lessons learned in addressing adaptation in the least developed countries through the national adaptation programme of action process. Volume 1. United Nations Climate Change Secretariat, Bonn, pp. 78. Leary, N., J. Adejuwon, V. Barros, P. Batimaa, B. Biagini, I. Burton, S. Chinvanno, R. Cruz, D. Dabi, A. de Comarmond, B. Dougherty, P. Dube, A. Githeko, A. Abou Hadid, M. Hellmuth, R. Kangalawe, J. Kulkarni, M. Kumar, R. Lasco, M. Mataki, M. Medany, M. Mohsen, G. Nagy, M. Njie, J. Nkomo, A. Nyong, B. Osman, E. Sanjak, R. Seiler, M. Taylor, M. Travasso, G. von Maltitz, S. Wandiga, and M. Wehbe, 2008a: A Stitch in time: general lessons from specific cases. In: Climate Change and Adaptation. [N. Leary, J. Adejuwon, V. Barros, I. Burton, J. Kulkarni and R. Lasco (eds.)]. Earthscan, London, UK and Sterling, VA, USA, pp. 1-27. Leary, N., J. Adejuwon, V. Barros, I. Burton, J. Kulkarni, and R. Lasco (eds.), 2008b: Climate change and adaptation Earthscan, London, UK and Sterling, VA, USA, pp. 398. Lebel, T. and A. Ali, 2009: Recent trends in the Central and Western Sahel rainfall regime (1990-2007). Journal of Hydrology, 375(1-2), 52-64. Leck, H., C. Sutherland, D. Scott, and G. Oelofse, 2012: Social and Cultural Barriers to Adaptation Implementation: the case of South Africa. In: Overcoming Barriers to Climate Change Adaptation Implementation in Southern Africa. [Masters, L. and L. Duff (eds.)]. Africa Institute of South Africa, Pretoria, South Africa, pp. 61-82. Lecocq, F. and Z. Shalizi, 2007: How Might Climate Change Affect Economic Growth in Developing Countries? A Review of the Growth Literature with a Climate Lens. In: Policy Research Working Paper 4315. The World Bank, pp. 54. Lericollais, A., 1989: La mort des arbres a Sob, en pays Sereer (Sénégal). In: Tropiques: Lieux et Liens. [Benoît, A., C. Blanc-Pamard, J.L. Chaléard, A. Dubresson, V. Lassailly-Jacob, J.M. Marchal et al. (eds.)]. Éditions ORSTOM, Paris, France, pp. 187-197. Levine, S., E. Ludi, and L. Jones, 2011: Rethinking support for adaptive capacity to climate change: the role of development interventions. Findings from Mozambique, Uganda and Ethiopia. A report for the Africa Climate Change Resilience Alliance. Overseas Development Institute (ODI), London, UK, pp. 38. Levine, S., 2012: The 2007/2008 food price acceleration in Namibia: an overview of impacts and policy responses. Food Security, 4(1), 59-71. Lewis, S.L., G. Lopez-Gonzalez, B. Sonké, K. Affum-Baffoe, T.R. Baker, L.O. Ojo, O.L. Phillips, J.M. Reitsma, L. White, J.A. Comiskey, M.-N.K. Djuikouo, C.E.N. Ewango, T.R. Feldpausch, A.C. Hamilton, M. Gloor, T. Hart, A. Hladik, J. Lloyd, J.C. Lovett, J.-R. Makana, Y. Malhi, F.M. Mbago, H.J. Ndangalasi, J. Peacock, K.S.-H. Peh, D. Sheil, T. Sunderland, M.D. Swaine, J. Taplin, D. Taylor, S.C. Thomas, R. Votere, and H. Wöll, 2009: Increasing carbon storage in intact African tropical forests. Nature, 457(7232), 1003-1006. Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Liebmann, B., I. Bladé, G.N. Kiladis, L.M.V. Carvalho, G.B. Senay, D. Allured, S. Leroux, and C. Funk, 2012: Seasonality of African precipitation from 1996 to 2009. Journal of Climate, 25(12), 4304-4322. Liller, H.B. and K. Van den Broeck, 2011: Economic drivers of migration and climate change in LDCs. Global Environmental Change, 21(SUPPL. 1), S70-S81. Linnerooth-Bayer, J., K. Warner, C. Bals, P. Höppe, I. Burton, T. Loster, and A. Haas, 2009: Insurance, developing countries and climate change. The Geneva Papers on Risk and Insurance - Issues and Practice, 34, 381-400. Linthicum, K.J., A. Anyamba, C.J. Tucker, P.W. Kelley, M.F. Myers, and C.J. Peters, 1999: Climate and satellite indicators to forecast Rift Valley fever epidemics in Kenya. Science, 285(5426), 397-400. Lionello, P. and F. Giorgi, 2007: Winter precipitation and cyclones in the Mediterranean region: Future climate scenarios in a regional simulation. Advances in Geosciences, 12, 153-158. Liu, J., S. Fritz, C.F.A. van Wesenbeeck, M. Fuchs, L. You, M. Obersteiner, and H. Yang, 2008: A spatially explicit assessment of current and future hotspots of hunger in Sub-Saharan Africa in the context of global change. Global and Planetary Change, 64(3-4), 222-235. Liu, L., H.L. Johnson, S. Cousens, J. Perin, S. Scott, J.E. Lawn, I. Rudan, H. Campbell, R. Cibulskis, M. Li, C. Mathers, and R.E. Black, 2012: Global, regional, and national causes of child mortality: An updated systematic analysis for 2010 with time trends since 2000. The Lancet, 379(9832), 2151-2161. Lloyd, S.J., R. Sari Kovats, and Z. Chalabi, 2011: Climate change, crop yields, and undernutrition: Development of a model to quantify the impact of climate scenarios on child undernutrition. Environmental Health Perspectives, 119(12), 1817-1823. Lobell, D.B., M. Bänziger, C. Magorokosho, and B. Vivek, 2011: Nonlinear heat effects on African maize as evidenced by historical yield trials. Nature Climate Change, 1, 42-45. Lobell, D.B., M.B. Burke, C. Tebaldi, M.D. Mastrandrea, W.P. Falcon, and R.L. Naylor, 2008: Prioritizing climate change adaptation needs for food security in 2030. Science, 319(5863), 607-610. Lockwood, M., 2012: Adaptation policy, governance and politics in sub-Saharan Africa. Paper to be presented at an international symposium on the Governance of Adaptation, Amsterdam, March 22-23 2012. Institute of Development Studies (IDS), Sussex, UK, pp. 29. Loha, E. and B. Lindtjrn, 2010: Model variations in predicting incidence of Plasmodium falciparum malaria using 1998-2007 morbidity and meteorological data from south Ethiopia. Malaria Journal, 9, 166. Lou, Y. and X.-Q. Zhao, 2010: A climate-based malaria transmission model with structured vector population. SIAM Journal on Applied Mathematics, 70(6), 2023-2044. Ludi, E., L. Jones, and S. Levine, 2012: Changing focus? How to take adaptive capacity seriously. Evidence from Africa shows that development interventions could do more. ODI Briefing Paper 71. Overseas Development Institute (ODI), London, UK, pp. 4. Lunde, T.M., M.N. Bayoh, and B. Lindtjrn, 2013: How malaria models relate temperature to malaria transmission. Parasites and Vectors, 6(1), 20. Lunde, T.M. and B. Lindtjrn, 2013: Cattle and climate in Africa: How climate variability has influenced national cattle holdings from 1961-2008. PeerJ, 1, e55. Luque Fernández, M.A., A. Bauernfeind, J.D. Jiménez, C.L. Gil, N.E. Omeiri, and D.H. Guibert, 2009: Influence of temperature and rainfall on the evolution of cholera epidemics in Lusaka, Zambia, 2003-2006: analysis of a time series. Transactions of the Royal Society of Tropical Medicine and Hygiene, 103(2), 137-143. Lutz, W. and S. K.C., 2010: Dimensions of global population projections: What do we know about future population trends and structures? Philosophical Transactions of the Royal Society B: Biological Sciences, 365(1554), 2779-2791. Lwasa, S., 2010: Adapting urban areas in Africa to climate change: The case of Kampala. Current Opinion in Environmental Sustainability, 2(3), 166-171. Lyon, B., 2009: Southern Africa summer drought and heat waves: Observations and coupled model behavior. Journal of Climate, 22(22), 6033-6046. Lyon, B. and D.G. DeWitt, 2012: A recent and abrupt decline in the East African long rains. Geophysical Research Letters, 39(2). Lyon, B., A.G. Barnston, and D.G. DeWitt, 2013: Tropical pacific forcing of a 1998-1999 climate shift: observational analysis and climate model results for the boreal spring season. Climate Dynamics,(Published online 26-07-2013), 1-17. Lyons, C.L., M. Coetzee, J.S. Terblanche, and S.L. Chown, 2012: Thermal limits of wild and laboratory strains of two African malaria vector species, Anopheles arabiensis and Anopheles funestus. Malaria Journal, 11, 226. Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Lyons, C.L., M. Coetzee, and S.L. Chown, 2013: Stable and fluctuating temperature effects on the development rate and survival of two malaria vectors, Anopheles arabiensis and Anopheles funestus. Parasites and Vectors, 6(1), 104. Mabaso, M.L.H., I. Kleinschmidt, B. Sharp, and T. Smith, 2007: El Nino Southern Oscillation (ENSO) and annual malaria incidence in Southern Africa. Transactions of the Royal Society of Tropical Medicine and Hygiene, 101(4), 326-330. Macchi, M., G. Oviedo, S. Gotheil, K. Cross, A. Boedhihartono, C. Wolfangel, and M. Howell, 2008: Indigenous and traditional peoples and climate change. IUCN Issues Paper. International Union for Conservation of Nature (IUCN), Gland, Switzerland, pp. 64. MacDonald, A.M., R.C. Carlow, D.M.J. MacDonald, W.G. Darling, and B.ÉÓ Dochartaigh, 2009: What impact will climate change have on rural groundwater supplies in Africa? Hydrological Sciences Journal, 54(4), 690-703. MacDonald, A.M., H.C. Bonsor, B.ÉÓ Dochartaigh, and R.G. Taylor, 2012: Quantitative maps of groundwater resources in Africa. Environmental Research Letters, 7(2). Madzwamuse, M., 2010: Drowning voices: the climate change discourse in South Africa. Policy Brief 05/2010. Heinrich-Böll-Foundation Southern Africa, Cape Town, South Africa, pp. 8. Magadza, C.H.D., 2008: Limnology News from Zimbabwe. SIL News, 52, 12 October 2009. Magadza, C.H.D., 2010: Environmental state of Lake Kariba and Zambezi River Valley: Lessons learned and not learned. Lakes and Reservoirs: Research and Management, 15(3), 167-192. Magadza, C.H.D., 2011: Indications of the effects of climate change on the pelagic fishery of Lake Kariba, Zambia- Zimbabwe. Lakes and Reservoirs: Research and Management, 16(1), 15-22. Malhi, Y., S. Adu-Bredu, R.A. Asare, S.L. Lewis, and P. Mayaux, 2013: African rainforests: Past, present and future. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1625). Mamouda, M.N.A., 2011: Africa's National Adaptation Programmes of Action. In: Joto Africa - adapting to climate change in Africa [Denton, F. (ed.)]. Issue 8. Arid Lands Information Network, Nairobi, Kenya, pp. 8. Mandleni, B. and F.D.K. Anim, 2011: Perceptions of Cattle and Sheep Farmers on Climate Change and Adaptation in the Eastern Cape Province of South Africa. Journal of Human Ecology, 34(2), 107-112. Marchiori, L., J.-F. Maystadt, and I. Schumacher, 2011: The impact of climate variations on migration in sub- Saharan Africa. Paper presented at CSAE 25th Anniversary Conference 2011: Economic Development in Africa, Oxford, UK, pp. 33. Mariotti, L., E. Coppola, M.B. Sylla, F. Giorgi, and C. Piani, 2011: Regional climate model simulation of projected 21st century climate change over an all-Africa domain: Comparison analysis of nested and driving model results. Journal of Geophysical Research D: Atmospheres, 116(15). Marongwe, L.S., K. Kwazira, M. Jenrich, C. Thierfelder, A. Kassam, and T. Friedrich, 2011: An African success: The case of conservation agriculture in Zimbabwe. International Journal of Agricultural Sustainability, 9(1), 153-161. Marshall, B., C. Ezekiel, J. Gichuki, O. Mkumbo, L. Sitoki, and F. Wanda, 2009: Global warming is reducing thermal stability and mitigating the effects of eutrophication in Lake Victoria (East Africa). Nature Precedings,. Marshall, B.E., 2012: Does climate change really explain changes in the fisheries productivity of Lake Kariba (Zambia-Zimbabwe)? Transactions of the Royal Society of South Africa, 67(1), 45-51. Martens, P., D. McEvoy, and C. Chang, 2009: The climate change challenge: linking vulnerability, adaptation, and mitigation. Current Opinion in Environmental Sustainability, 1(1), 14-18. Masike, S. and P.B. Urich, 2009: The projected cost of climate change to livestock water supply and implications in Kgatleng District, Botswana. World Journal of Agricultural Sciences, 5(5), 597-603. Masike, S. and P. Urich, 2008: Vulnerability of traditional beef sector to drought and the challenges of climate change: The case of Kgatleng District, Botswana. Journal of Geography and Regional Planning, 1(1), 12-18. Mason, N.M., T.S. Jayne, A. Chapoto, and C. Donovan, 2011: Putting the 2007/2008 global food crisis in longer- term perspective: Trends in staple food affordability in urban Zambia and Kenya. Food Policy, 36(3), 350-367. Mason, P.R., 2009: Zimbabwe experiences the worst epidemic of cholera in Africa. Journal of Infection in Developing Countries, 3(2), 148-151. Mathee, A., J. Oba, and A. Rose, 2010: Climate change impacts on working people (the HOTHAPS initiative): findings of the South African pilot study. Global Health Action, 3, 5612. Mather, A.A. and D.D. Stretch, 2012: A perspective on sea level rise and coastal storm surge from Southern and Eastern Africa: A case study near Durban, South Africa. Water, 4, 237-259. Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mayaux, P., J.-F. Pekel, B. Desclée, F. Donnay, A. Lupi, F. Achard, M. Clerici, C. Bodart, A. Brink, R. Nasi, and A. Belward, 2013: State and evolution of the African rainforests between 1990 and 2010. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1625). McCartney, M.P. and M. Menker Girma, 2012: Evaluating the downstream implications of planned water resource development in the Ethiopian portion of the Blue Nile River. Water International, 37(4), 362-379. McClanahan, T.R., N.A. Muthiga, J. Maina, A.T. Kamukuru, and S.A.S. Yahya, 2009: Changes in northern Tanzania coral reefs during a period of increased fisheries management and climatic disturbance. Aquatic Conservation: Marine and Freshwater Ecosystems, 19(7), 758-771. McCully, P., 2007: Before the Deluge: Coping with floods in a changing climate. IRN Dams, Rivers and People Report 2007. International River Network, Berkeley, CA, USA, pp. 24. McFerson, H.M., 2010: Poverty among women in Sub-Saharan Africa: A review of selected issues. Journal of International Women's Studies, 11(4), 50-72. Meehl, G.A., C. Covey, T. Delworth, M. Latif, B. McAvaney, J.F.B. Mitchell, R.J. Stouffer, and K.E. Taylor, 2007: The WCRP CMIP3 multimodel dataset: A new era in climatic change research. Bulletin of the American Meteorological Society, 88(9), 1383-1394. Mendelsohn, J. and T. Dawson, 2008: Climate and cholera in KwaZulu-Natal, South Africa: The role of environmental factors and implications for epidemic preparedness. International Journal of Hygiene and Environmental Health, 211(1-2), 156-162. Mertz, O., C. Mbow, A. Reenberg, and A. Diouf, 2009: Farmers' perceptions of climate change and agricultural adaptation strategies in rural sahel. Environmental Management, 43(5), 804-816. Mertz, O., C. Mbow, A. Reenberg, L. Genesio, E.F. Lambin, S. D'haen, M. Zorom, K. Rasmussen, D. Diallo, B. Barbier, I.B. Moussa, A. Diouf, J.O Nielsen, and I. Sandholt, 2011: Adaptation strategies and climate vulnerability in the Sudano-Sahelian region of West Africa. Atmospheric Science Letters, 12(1), 104-108. Midgley, G., 2013: Long Term Adaptation Scenarios Research Flagship Programme (LTAS): Phase 1. Technical Report . Technical Summary. South African National Biodiversity Institute, pp. 79. Milder, J.C., T. Majanen, and S.J. Scherr, 2011: Performance and potential of conservation agriculture for climate change adaptation and mitigation in Sub-Saharan Africa . Ecoagriculture Discussion Paper no. 6. Ecoagriculture Partners, Washington, DC, USA, pp. 91. Milzow, C., V. Burg, and W. Kinzelbach, 2010: Estimating future ecoregion distributions within the Okavango Delta Wetlands based on hydrological simulations and future climate and development scenarios. Journal of Hydrology, 381(1-2), 89-100. Min, S.-K. and A. Hense, 2007: A Bayesian assessment of climate change using multimodel ensembles. Part II: Regional and seasonal mean surface temperatures. Journal of Climate, 20(12), 2769-2790. Mitchard, E.T.A. and C.M. Flintrop, 2013: Woody encroachment and forest degradation in sub-Saharan Africa's woodlands and savannas 1982-2006. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1625). Moench, M., 2011: Aren t we all vulnerable: why do vulnerability analysis?. IIED Opinion: Lessons from adaptation in practice, November 2011. International Institute for Environment and Development (IIED), London, UK, pp. 2. Mogoi, J.A., E.K. Obonyo, J. Tanui, D. Catacutan, and J.G. Mowo, 2010: Strengthening local institutions as avenues for climate change resilience. pp. 22. Mohamed, A.B., 2011: Climate change risks in Sahelian Africa. Regional Environmental Change, 11(SUPPL. 1), 109-117. Mohino, E., S. Janicot, and J. Bader, 2011: Sahel rainfall and decadal to multi-decadal sea surface temperature variability. Climate Dynamics, 37(3), 419-440. Moise, A.F. and D.A. Hudson, 2008: Probabilistic predictions of climate change for Australia and southern Africa using the reliability ensemble average of IPCC CMIP3 model simulations. Journal of Geophysical Research D: Atmospheres, 113(15). Molesworth, A.M., L.E. Cuevas, S.J. Connor, A.P. Morse, and M.C. Thomson, 2003: Environmental risk and meningitis epidemics in Africa. Emerging Infectious Diseases, 9(10), 1287-1293. Monerie, P.-A., B. Fontaine, and P. Roucou, 2012: Expected future changes in the African monsoon between 2030 and 2070 using some CMIP3 and CMIP5 models under a medium-low RCP scenario. Journal of Geophysical Research D: Atmospheres, 117(16). Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mora, C.F.,A.G., R.J. Longman, R.S. Dacks, M.M. Walton, E.J. Tong, J.J. Sanchez, L.R. Kaiser, Y.O. Stender, J.M. Anderson, C.M. Ambrosino, I. Fernandez-Silva, L.M. Giuseffi, and T.W. Giambelluca, 2013: The projected timing of climate departure from recent variability. Nature, 502, 183-187. Mordecai, E.A., K.P. Paaijmans, L.R. Johnson, C. Balzer, T. Ben-Horin, E. de Moor, A. Mcnally, S. Pawar, S.J. Ryan, T.C. Smith, and K.D. Lafferty, 2013: Optimal temperature for malaria transmission is dramatically lower than previously predicted. Ecology Letters, 16(1), 22-30. Mortimore, M., 2010: Adapting to drought in the Sahel: Lessons for climate change. Wiley Interdisciplinary Reviews: Climate Change, 1(1), 134-143. Moser, C. and D. Satterthwaite, 2008: Towards Pro-Poor Adaptation to Climate Change in the Urban Centres of Low- and Middle- Income Countries. Global Urban Research Centre Working Paper #1. The Global Urban Research Centre (GURC), University of Manchester, Manchester, UK, pp. 44. Moser, S.C. and J.A. Ekstrom, 2010: A framework to diagnose barriers to climate change adaptation. Proceedings of the National Academy of Sciences of the United States of America, 107(51), 22026-22031. Mosha, A.C., 2011: The Effects of Climate Change on Urban Human Settlements in Africa. In: Climate Change and Sustainable Urban Development in Africa and Asia. [Yuen, B. and A. Kumssa (eds.)]. Springer, Dordrecht, Heidelberg, London, New York, pp. 69-100. Moss, B., 2010: Climate change, nutrient pollution and the bargain of Dr Faustus. Freshwater Biology, 55(SUPPL. 1), 175-187. Motsholapheko, M.R., D.L. Kgathi, and C. Vanderpost, 2011: Rural livelihoods and household adaptation to extreme flooding in the Okavango Delta, Botswana. Physics and Chemistry of the Earth, 36(14-15), 984-995. Mougou, R., M. Mansour, A. Iglesias, R.Z. Chebbi, and A. Battaglini, 2011: Climate change and agricultural vulnerability: A case study of rain-fed wheat in Kairouan, Central Tunisia. Regional Environmental Change, 11(SUPPL. 1), 137-142. Moustadraf, J., M. Razack, and M. Sinan, 2008: Evaluation of the impacts of climate changes on the coastal Chaouia aquifer, Morocco, using numerical modeling. Hydrogeology Journal, 16(7), 1411-1426. Moyo, D., 2009: Dead Aid: Why aid is not working and how there is a better way for Africa. Penguin Books, London, pp. 208. Mrema, S., A. Shamte, M. Selemani, and H. Masanja, 2012: The influence of weather on mortality in rural Tanzania: A time-series analysis 1999-2010. Global Health Action, 5(SUPPL.), 33-43. Mubaya, C.P., J. Njuki, E.P. Mutsvangwa, F.T. Mugabe, and D. Nanja, 2012: Climate variability and change or multiple stressors? Farmer perceptions regarding threats to livelihoods in Zimbabwe and Zambia. Journal of Environmental Management, 102, 9-17. Muchena, F.N., D.D. Onduru, G.N. Gachini, and A. de Jager, 2005: Turning the tides of soil degradation in Africa: Capturing the reality and exploring opportunities. Land use Policy, 22(1), 23-31. Mukheibir, P. and G. Ziervogel, 2007: Developing a Municipal Adaptation Plan (MAP) for climate change: The city of Cape Town. Environment and Urbanization, 19(1), 143-158. Mukute, M., T. Marange, C. Masara, H. Lotz-Sisitka, and T. Pesanayi, 2012: Assessment of Southern African Development Community (SADC) Capacity to Implement International Agreements on Environment, Sustainable Development and Education. In: Promoting critical knowledge, skills and qualifications for sustainable development in Africa: How to design and implement an effective response through education and training systems. Working Document 3.3.03 for the Triennale on Education and Training in Africa (Ouagadougou, Burkina Faso, February 7 - February 12, 2012). Association for the Development of Education in Africa (ADEA), pp. 58. Murray, V., G. McBean, M. Bhatt, S. Borsch, T.S. Cheong, W.F. Erian, S. Llosa, F. Nadim, M. Nunez, R. Oyun, and A.G. Suarez, 2012: Case studies. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX). Special Report of the Intergovernmental Panel on Climate Change (IPCC). [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al. (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 487-542. Murray-Hudson, M., P. Wolski, and S. Ringrose, 2006: Scenarios of the impact of local and upstream changes in climate and water use on hydro-ecology in the Okavango Delta, Botswana. Journal of Hydrology, 331(1-2), 73- 84. Mutonyi, S. and B. Fungo, 2011: Patterns of agroforestry practices among small-holder farmers in the Lake Victoria Crescent Zone (LVCAEZ) of Uganda. Research Journal of Applied Sciences, 6(4), 251-257. Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mutuku, F.M., C.H. King, A.L. Bustinduy, P.L. Mungai, E.M. Muchiri, and U. Kitron, 2011: Impact of drought on the spatial pattern of transmission of Schistosoma haematobium in coastal Kenya. American Journal of Tropical Medicine and Hygiene, 85(6), 1065-1070. Mwaura, P., 2008: Indigenous Knowledge in Disaster Management in Africa. United Nations Environment Programme (UNEP), Nairobi, Kenya, pp. 117. Naess, L.O., E. Polack, and B. Chinsinga, 2011: Bridging Research and Policy Processes for Climate Change Adaptation. IDS Bulletin, 42(3), 97-103. Naidoo, T., K. Vaz, and L. Byaba, 2012: An introduction to adaptation in view of developing a private financing approach for adaptation related projects using the CTI PFAN model. CTI PFAN Background Paper on Adaptation 02/2012. Climate Technology Initiative, pp. 51. Naidu, S., R. Hounsome, and K. Iyer, 2006: Climatic futures for Durban: revised report. [Roberts, D., Mather, A. and Maganlal, M. (eds.)]. CSIR Environmentek and Ethekwini Municipality, pp. 222. Nakaegawa, T., C. Wachana, and KAKUSHIN Team-3 Modeling Group, 2012: First impact assessment of hydrological cycle in the Tana River Basin, Kenya, under a changing climate in the late 21st Century. Hydrological Research Letters, 6, 29-34. Nakhooda, S., A. Caravani, N. Bird, and L. Schalatek, 2011: Climate Finance in Sub-Saharan Africa. Climate Finance Policy Briefs. Heinrich Böll Stiftung North America and Overseas Development Institute, pp. 8. Narayan, N., A. Paul, S. Mulitza, and M. Schulz, 2010: Trends in coastal upwelling intensity during the late 20th century. Ocean Science, 6(3), 815-823. Nasi, R., D. Brown, D. Wilkie, E. Bennett, C. Tutin, G. van Tol, and T. Christophersen, 2008: Conservation and use of wildlife-based resources: the bushmeat crisis. Technical Series no. 33. Secretariat of the Convention on Biological Diversity (CBD) and Center for International Forestry Research (CIFOR), Montreal, Canada and Bogor, Indonesia, pp. 50. Naude, W., 2010: Environment, Forced Migration and Social Vulnerability. In: Forced Migration from Sub-Saharan Africa: The Conflict Environment Link [Afifi, T. and J. Jäger(eds.)]. Springer, Berlin, Heidelberg, pp. 43-55. Ncube, S.P., H. Makurira, E. Kaseke, and A. Mhizha, 2011: Reservoir operation under variable climate: Case of Rozva Dam, Zimbabwe. Physics and Chemistry of the Earth, 36(14-15), 1112-1119. Ndebele-Murisa, M.R., 2011: An analysis of primary and secondary production in Lake Kariba in a changing climate. Diss. PhD, University of the Western Cape, South Africa, 181 pp. Ndebele-Murisa, M.R., M.C. Frank, and R.L. Miles, 2012: Phytoplankton biomass and primary production dynamics in Lake Kariba. Lakes and Reservoirs: Research and Management, 17(4), 275-289. Ndebele-Murisa, M.R., E. Mashonjowa, and T. Hill, 2011: The implications of a changing climate on the Kapenta fish stocks of Lake Kariba, Zimbabwe. Transactions of the Royal Society of South Africa, 66(2), 105-119. Ndebele-Murisa, M.R., C.F. Musil, and L. Raitt, 2010: A review of phytoplankton dynamics in tropical African lakes. South African Journal of Science, 106(1-2), 13-18. Ndegwa, W., K.P.C. Rao, R.K. Ngugi, and K. Kwena, 2010: Improving farmer adaptive capacity by integrating local and indigenous knowledge in climate forecasting and adaptive response. Adaptation Insights No. 4. IDRC anda CIFOR, pp. 3. Nelson, G.C., M.W. Rosegrant, J. Koo, R. Robertson, T. Sulser, T. Zhu, C. Ringler, S. Msangi, A. Palazzo, M. Batka, M. Magalhaes, R. Valmonte-Santos, M. Ewing, and D. Lee, 2009: Climate change: Impact on agriculture and costs of adaptation. Food Policy Report #19. International Food Policy Research Institute (IFPRI), Washington, D.C., pp. 30. Nelson, F. and A. Agrawal, 2008: Patronage or participation? Community-based natural resource management reform in Sub-Saharan Africa. Development and Change, 39(4), 557-585. NEPAD, 2010: CAADP - Highlighting the successes. The NEPAD Planning and Coordinating Agency (NEPAD Agency), Addis Ababa, Ethiopia, pp. 39. NEPAD, ECA, and OSAA, 2012: Africa's decade of change: reflections on 10 years of NEPAD. NEPAD Planning and Coordinating Agency, Economic Commission for Africa (ECA) and Office of the Special Adviser on Africa (OSAA), Addis Ababa, Ethiopia, pp. 48. Neumann, R., G. Jung, P. Laux, and H. Kunstmann, 2007: Climate trends of temperature, precipitation and river discharge in the Volta Basin of West Africa. International Journal of River Basin Management, 5(1), 17-30. New, M., B. Hewitson, D.B. Stephenson, A. Tsiga, A. Kruger, A. Manhique, B. Gomez, C.A.S. Coelho, D.N. Masisi, E. Kululanga, E. Mbambalala, F. Adesina, H. Saleh, J. Kanyanga, J. Adosi, L. Bulane, L. Fortunata, Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 M.L. Mdoka, and R. Lajoie, 2006: Evidence of trends in daily climate extremes over southern and west Africa. Journal of Geophysical Research D: Atmospheres, 111(14). Newsham, A.J. and D.S.G. Thomas, 2011: Knowing, farming and climate change adaptation in North-Central Namibia. Global Environmental Change, 21(2), 761-770. Ngigi, S.N., J. Rockström, and H.H.G. Savenije, 2006: Assessment of rainwater retention in agricultural land and crop yield increase due to conservation tillage in Ewaso Ng'iro river basin, Kenya. Physics and Chemistry of the Earth, 31(15-16), 910-918. Nguku, P., S. Sharif, A. Omar, C. Nzioka, P. Muthoka, J. Njau, A. Dahiye, T. Galgalo, J. Mwihia, J. Njoroge, H. Limo, J. Mutiso, R. Kalani, A. Sheikh, J. Nyikal, D. Mutonga, J. Omollo, A. Guracha, J. Muindi, S. Amwayi, D. Langat, D. Owiti, A. Mohammed, J. Musaa, R. Sang, R. Breiman, K. Njenga, D. Feikin, M. Katz, H. Burke, P. Nyaga, M. Ackers, S. Gikundi, V. Omballa, L. Nderitu, N. Wamola, R. Wanjala, S. Omulo, J. Richardson, D. Schnabel, S. Martin, D. Hoel, H. Hanafi, M. Weiner, J. Onsongo, T. Kojo, M. Duale, A. Hassan, M. Dabaar, C. Njuguna, M. Yao, T. Grein, P. Formenty, B. Telfer, R. Lepec, H. Feldmann, A. Grolla, S. Wainwright, E. Lederman, E. Farnon, C. Rao, B.K. Kapella, and H. Gould, 2007: Rift Valley fever outbreak - Kenya, November 2006-January 2007. Morbidity and Mortality Weekly Report, 56(4), 73-76. Ngwira, A.R., J.B. Aune, and S. Mkwinda, 2012: On-farm evaluation of yield and economic benefit of short term maize legume intercropping systems under conservation agriculture in Malawi. Field Crops Research, 132, 149-157. Nhemachena, C. and R.M. Hassan, 2011: Micro-Level Analysis of Farmers Adaptation to Climate Change in Southern Africa. In: How can African agriculture adapt to climate change? Insights from Ethiopia and South Africa [Ringler, C., Bryan, E., Hassan, R.M., Alemu, T. and Hillesland, M. (eds.)]. IFPRI Research Brief No. 15. International Food Policy Research Institute (IFPRI), Washington, DC, USA, pp. 2. Niang, I., 2007: Institutional Framework in Relation to Climate Change in West and Central Africa. Climate Change Adaptation in Africa Programme, Dakar, pp. 49. Niang, I., M. Dansokho, S. Faye, K. Gueye, and P. Ndiaye, 2010: Impacts of climate change on the Senegalese coastal zones: Examples of the Cap Vert peninsula and Saloum estuary. Global and Planetary Change, 72(4), 294-301. Niang, I., 2012: Coastal erosion and the adaptation to climate change in coastal zones of West Africa project. ACCC Project Brief. Adaptation to Climate Change in Coastal Zones of West Africa (ACCC), pp. 249-250. Nicholls, T., L. Norgrove, and G. Masters, 2008: Innovative solutions to new invaders: managing agricultural pests, diseases and weeds under climate change[Brown, A.G. (ed.)]. Proceedings of Agriculture in a Changing Climate. Records of a conference conducted by the ATSE Crawford Fund. Fourteenth Annual Development Conference, 3 September 2008, Parliament House, Canberra, Australia, pp. 9-14. Nicholson, S.E., B. Some, and B. Kone, 2000: An analysis of recent rainfall conditions in West Africa, including the rainy seasons of the 1997 El Nino and the 1998 La Nina years. Journal of Climate, 13(14), 2628-2640. Nicholson, S.E., D.J. Nash, B.M. Chase, S.W. Grab, T.M. Shanahan, D. Verschuren, A. Asrat, A.-M. Lézine, and M. Umer, 2013: Temperature variability over Africa during the last 2000 years. Holocene, 23(8), 1085-1094. Nielsen, J.O, 2010: The outburst: Climate change, gender relations, and situational analysis. Social Analysis, 54(3), 76-89. Nielsen, J.O and A. Reenberg, 2010: Cultural barriers to climate change adaptation: A case study from Northern Burkina Faso. Global Environmental Change, 20(1), 142-152. Nikulin, G., C. Jones, F. Giorgi, G. Asrar, M. Büchner, R. Cerezo-Mota, O.B. Christensen, M. Déqué, J. Fernandez, A. Hänsler, E. van Meijgaard, P. Samuelsson, M.B. Sylla, and L. Sushama, 2012: Precipitation climatology in an ensemble of CORDEX-Africa regional climate simulations. Journal of Climate, 25(18), 6057-6078. Nilsson, A. and A. Schnell, 2010: Lessons learnt on HRBA in Sub-Saharan Africa . pp. 10. Nin-Pratt, A., M. Johnson, and B. Yu, 2012: Improved performance of agriculture in Africa South of the Sahara: Taking off or bouncing back. IFPRI Discussion Paper 01224. International Food Policy Research Institute (IFPRI), Washington, DC, USA, pp. 32. Njie, M., B.E. Gomez, M.E. Hellmuth, J.M. Callaway, B.P. Jallow, and P. Droogers, 2008: Making economic sense of adaptation in upland cereal production systems in The Gambia. In: Climate change and vulnerability and adaptation. [Leary, N., J. Adejuwon, V. Barros, I. Burton, J. Kulkarni, and R. Lasco (eds.)]. Earthscan, London, UK and Sterling, VA, USA, pp. 131-146. Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Nkem, J., H. Santoso, D. Murdiyarso, M. Brockhaus, and M. Kanninen, 2007: Using tropical forest ecosystem goods and services for planning climate change adaptation with implications for food security and poverty reduction. SAT eJournal, 4(1), 1-23. Nkem, J., M. Idinoba, M. Brockhaus, F. Kalame, and A. Tas, 2008: Adaptation to climate change in Africa: Synergies with biodiversity and forest. CIFOR Environment Brief No. 3. Center for International Forestry Research (CIFOR), Bogor, Indonesia, pp. 4. Notter, B., H. Hurni, U. Wiesmann, and J.O. Ngana, 2012: Evaluating watershed service availability under future management and climate change scenarios in the Pangani Basin. Physics and Chemistry of the Earth, 61-62, 1- 11. Nyanga, P.H., F.H. Johnsen, J.B. Aune, and T.H. Kalinda, 2011: Smallholder Farmers Perceptions of Climate Change and Conservation Agriculture: Evidence from Zambia. Journal of Sustainable Development, 4(4), 73- 85. Nyenje, P.M. and O. Batelaan, 2009: Estimating the effects of climate change on groundwater recharge and baseflow in the upper Ssezibwa catchment, Uganda. Hydrological Sciences Journal, 54(4), 713-726. Nyong, A., F. Adesina, and B. Osman Elasha, 2007: The value of indigenous knowledge in climate change mitigation and adaptation strategies in the African Sahel. Mitigation and Adaptation Strategies for Global Change, 12(5), 787-797. Nyssen, J., J. Poesen, D. Gebremichael, K. Vancampenhout, M. D'aes, G. Yihdego, G. Govers, H. Leirs, J. Moeyersons, J. Naudts, N. Haregeweyn, M. Haile, and J. Deckers, 2007: Interdisciplinary on-site evaluation of stone bunds to control soil erosion on cropland in Northern Ethiopia. Soil and Tillage Research, 94(1), 151-163. Oates, N., D. Conway, and R. Calow, 2011: The mainstreaming approach to climate change adaptation: insights from Ethiopia s water sector. ODI Background Note. Overseas Development Institute (ODI), London, UK, pp. 8. Odada, E.O., D.O. Olago, and W.O. Ochola (eds.), 2006: Environment for Development: An ecosystems assessment of lake Victoria basin environmental and socio-economic status, trends and human vulnerabilities Pan African START Secretariat (PASS) and United Nations Environment Programme (UNEP), Nairobi, Kenya, pp. 192. Odongo-Aginya, E.I., F.K. Kironde, N.B. Kabatereine, P. Kategere, and F. Kazibwe, 2008: Effect of seasonal rainfall and other environmental changes, on snail density and infection rates with schistosoma mansoni fifteen years after the last snails' study in kigungu, entebbe, uganda. East African Medical Journal, 85(11), 556-563. Oguntunde, P.G. and B.J. Abiodun, 2013: The impact of climate change on the Niger River Basin hydroclimatology, West Africa. Climate Dynamics, 40(1-2), 81-94. Olago, D., M. Marshall, S.O. Wandiga, M. Opondo, P.Z. Yanda, R. Kanalawe, A.K. Githeko, T. Downs, A. Opere, R. Kavumvuli, E. Kirumira, L. Ogallo, P. Mugambi, E. Apindi, F. Githui, J. Kathuri, L. Olaka, R. Sigalla, R. Nanyunja, T. Baguma, and P. Achola, 2007: Climatic, socio-economic, and health factors affecting human vulnerability to cholera in the Lake Victoria basin, east Africa. Ambio, 36(4), 350-358. Olaka, L.A., E.O. Odada, M.H. Trauth, and D.O. Olago, 2010: The sensitivity of East African rift lakes to climate fluctuations. Journal of Paleolimnology, 44(2), 629-644. Olsson, P., L.H. Gunderson, S.R. Carpenter, P. Ryan, L. Lebel, C. Folke, and C.S. Holling, 2006: Shooting the rapids: Navigating transitions to adaptive governance of social-ecological systems. Ecology and Society, 11(1), 18. Olwoch, J.M., A.S. Van Jaarsveld, C.H. Scholtz, and I.G. Horak, 2007: Climate change and the genus rhipicephalus (Acari: Ixodidae) in Africa. Onderstepoort Journal of Veterinary Research, 74(1), 45-72. Olwoch, J.M., B. Reyers, F.A. Engelbrecht, and B.F.N. Erasmus, 2008: Climate change and the tick-borne disease, Theileriosis (East Coast fever) in sub-Saharan Africa. Journal of Arid Environments, 72(2), 108-120. Omumbo, J.A., B. Lyon, S.M. Waweru, S.J. Connor, and M.C. Thomson, 2011: Raised temperatures over the Kericho tea estates: Revisiting the climate in the East African highlands malaria debate. Malaria Journal, 10, 12. Orlove, B., C. Roncoli, M. Kabugo, and A. Majugu, 2010: Indigenous climate knowledge in southern Uganda: The multiple components of a dynamic regional system. Climatic Change, 100(2), 243-265. Orlowsky, B. and S.I. Seneviratne, 2012: Global changes in extreme events: Regional and seasonal dimension. Climatic Change, 110(3-4), 669-696. Osbahr, H., E. Boyd, and J. Martin, 2007: Resilience, Realities and Research in African Environments. Report of Workshop 18 June 2007. University of Oxford, Oxford, UK, pp. 24. Subject to Final Copyedit 85 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Osbahr, H., C. Twyman, W.N. Adger, and D.S.G. Thomas, 2010: Evaluating Successful Livelihood Adaptation to Climate Variability and change in Southern Africa. Ecology and Society, 15(2), 27. Osborne, P.L., 2012: Tropical ecosystems and ecological concept. Cambridge University Press, Cambridge, UK, Second ed., pp. 536. Osman-Elasha, B., N.Goutbi, E. Spanger-Siegfried, B. Dougherty, A. Hanafi, S. Zakieldeen, E. Sanjak, H. Atti and H. Elhassan, 2008: Community development and coping with drought in rural Sudan. In: Climate Change and Adaptation. [N. Leary, J. Adejuwon, V. Barros, I. Burton, J. Kulkarni and R. Lasco (eds.)]. Earthscan, London, UK and Sterling, VA, USA, pp. 90-108. Ototo, E.N., A.K. Githeko, C.L. Wanjala, and T.W. Scott, 2011: Surveillance of vector populations and malaria transmission during the 2009/10 El Nino event in the western Kenya highlands: Opportunities for early detection of malaria hyper-transmission. Parasites and Vectors, 4(1), 144. Otto, F.E.L., R.G. Jones, K. Halladay, and M.R. Allen, 2013: Attribution of changes in precipitation patterns in African rainforests. Philosophical Transactions of the Royal Society B: Biological Sciences, 368(1625). Paaijmans, K.P., A.F. Read, and M.B. Thomas, 2009: Understanding the link between malaria risk and climate. Proceedings of the National Academy of Sciences of the United States of America, 106(33), 13844-13849. Paaijmans, K.P., S. Blanford, A.S. Bell, J.I. Blanford, A.F. Read, and M.B. Thomas, 2010a: Influence of climate on malaria transmission depends on daily temperature variation. Proceedings of the National Academy of Sciences of the United States of America, 107(34), 15135-15139. Paaijmans, K.P., S.S. Imbahale, M.B. Thomas, and W. Takken, 2010b: Relevant microclimate for determining the development rate of malaria mosquitoes and possible implications of climate change. Malaria Journal, 9, 196. Paaijmans, K.P., S. Blanford, B.H.K. Chan, and M.B. Thomas, 2012: Warmer temperatures reduce the vectorial capacity of malaria mosquitoes. Biology Letters, 8(3), 465-468. Paavola, J., 2008: Livelihoods, vulnerability and adaptation to climate change in Morogoro, Tanzania. Environmental Science and Policy, 11(7), 642-654. Paeth, H., K. Born, R. Girmes, R. Podzun, and D. Jacob, 2009: Regional climate change in tropical and Northern Africa due to greenhouse forcing and land use changes. Journal of Climate, 22(1), 114-132. Pahl-Wostl, C., 2009: A conceptual framework for analysing adaptive capacity and multi-level learning processes in resource governance regimes. Global Environmental Change, 19(3), 354-365. Palmgren, H., 2009: Meningococcal disease and climate. Global Health Action, 2, 8. Pantuliano, S. and M. Wekesa, 2008: Improving drought response in pastoral areas of Ethiopia: Somali and Afar regions and Borena Zone of Oromiya region. Overseas Development Institute (ODI) - Humanitarian Policy Group, London, UK, pp. 39. Parfitt, J., M. Barthel, and S. MacNaughton, 2010: Food waste within food supply chains: Quantification and potential for change to 2050. Philosophical Transactions of the Royal Society B: Biological Sciences, 365(1554), 3065-3081. Parham, P.E. and E. Michael, 2010: Modeling the effects of weather and climate change on malaria transmission. Environmental Health Perspectives, 118(5), 620-626. Parham, P.E., D. Pople, C. Christiansen-Jucht, S. Lindsay, W. Hinsley, and E. Michael, 2012: Modeling the role of environmental variables on the population dynamics of the malaria vector Anopheles gambiae sensu stricto. Malaria Journal, 11, 271. Parry, M., A. Evans, M.W. Rosegrant, and T. Wheeler, 2009a: Climate change and hunger: responding to the challenge. World Food Programme, Rome, Italy, pp. 104. Parry, M., N. Arnell, P. Berry, D. Dodman, S. Fankhauser, C. Hope, S. Kovats, R. Nicholls, D. Satterthwaite, R. Tiffin, and T. Wheeler, 2009b: Assessing the costs of adaptation to climate change: A review of the UNFCCC and other recent estimates. International Institute for Environment and Development (IIED) and Grantham Institute for Climate Change, London, UK, pp. 111. Pascual, M., J.A. Ahumada, L.F. Chaves, X. Rodó, and M. Bouma, 2006: Malaria resurgence in the East African highlands: Temperature trends revisited. Proceedings of the National Academy of Sciences of the United States of America, 103(15), 5829-5834. Passioura, J., 2006: Increasing crop productivity when water is scarce - From breeding to field management. Agricultural Water Management, 80(1-3 SPEC. ISS.), 176-196. Patricola, C.M. and K.H. Cook, 2010: Northern African climate at the end of the twenty-first century: An integrated application of regional and global climate models. Climate Dynamics, 35(1), 193-212. Subject to Final Copyedit 86 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Patricola, C.M. and K.H. Cook, 2011: Sub-Saharan Northern African climate at the end of the twenty-first century: forcing factors and climate change processes. Climate Dynamics, 37(5-6), 1165-1188. Patt, A., A. Dazé, and P. Suarez, 2009: Gender and climate change vulnerability: what s the problem, what s the solution? In: Distributional Impacts of Climate Change and Disasters: Concepts and Cases. [Ruth, M. and M.E. Ibarraran (eds.)]. Edward Elgar, Cheltenham, UK, pp. 82-102. Patt, A.G. and D. Schröter, 2008: Perceptions of climate risk in Mozambique: Implications for the success of adaptation strategies. Global Environmental Change, 18(3), 458-467. Pauw, W.P., J. Ayers, and A. Kaere, 2012: The international architecture of climate adaptation finance and its implications for African institutions. In: Adapting institutions to meet Climate Change Impacts: can Africa meet the Challenges? [Denton, F. (ed.)]. IDRC/Springer. Pauw, W.P., 2013: The role of perception in subsistence farmer adaptation in Africa enriching the climate finance debate. International Journal of Climate Change Strategies and Management, 5(3), 267-284. Peach Brown, H.C., 2011: Gender, climate change and REDD+ in the Congo Basin forests of Central Africa. International Forestry Review, 13(2), 163-176. Peden, D., G. Taddesse, and A. Haileslassie, 2009: Livestock water productivity: Implications for sub-Saharan Africa. Rangeland Journal, 31(2), 187-193. Pelham, L., E. Clay, and T. Braunholz, 2011: Natural disasters: what is the role for social safety nets?. Social Protection Discussion Paper No. 1102. The World Bank, Washington, DC, USA, pp. 142. Perch-Nielsen, S.L., M.B. Bättig, and D. Imboden, 2008: Exploring the link between climate change and migration. Climatic Change, 91(3-4), 375-393. Persson, A, R.J.T. Klein, C. Kehler Siebert, A. Atteridge, B. Müller, J. Hoffmaister, M. Lazarus, and T. Takama, 2009: Adaptation finance under a Copenhagen agreed outcome. Research Report. Stockholm Environment Institute, Stockholm, Sweden, pp. 187. Peters, J. and S. Thielmann, 2008: Promoting biofuels: Implications for developing countries. Energy Policy, 36(4), 1538-1544. Petersen, C. and S. Holness, 2011: South Africa: Ecosystem-based planning for climate change. World Resources Report Case Study. World Resources Report, Washington, DC, USA, pp. 18. Peterson, A.T., 2009: Shifting suitability for malaria vectors across Africa with warming climates. BMC Infectious Diseases, 9, 59. Pettorelli, N., A.L.M. Chauvenet, J.P. Duffy, W.A. Cornforth, A. Meillere, and J.E.M. Baillie, 2012: Tracking the effect of climate change on ecosystem functioning using protected areas: Africa as a case study. Ecological Indicators, 20, 269-276. Phoon, S.Y., A.Y. Shamseldin, and K. Vairavamoorthy, 2004: Assessing impacts of climate change on Lake Victoria Basin, Africa. Proceedings of 30th WEDC International Conference: People-centred approaches to water and environmental sanitation, Vientiane, Lao PDR, pp. 392-397. Piguet, É, A. Pécoud, and P. de Guchteneire (eds.), 2011: Migration and Climate Change Cambridge University Press and UNESCO Publishing, New York, NY, USA, pp. 464. Piguet, E., 2010: Linking climate change, environmental degradation, and migration: A methodological overview. Wiley Interdisciplinary Reviews: Climate Change, 1(4), 517-524. Plummer, R., 2012: Fostering governance and building capacity for climate change adaptation: can adaptive co- management help?. pp. 18. Postigo, J.A.R., 2010: Leishmaniasis in the World Health Organization Eastern Mediterranean Region. International Journal of Antimicrobial Agents, 36(SUPPL. 1), S62-S65. Potapov, P.V., S.A. Turubanova, M.C. Hansen, B. Adusei, M. Broich, A. Altstatt, L. Mane, and C.O. Justice, 2012: Quantifying forest cover loss in Democratic Republic of the Congo, 2000-2010, with Landsat ETM+ data. Remote Sensing of Environment, 122, 106-116. Pramova, E., B. Locatelli, M. Brockhaus, and S. Fohlmeister, 2012: Ecosystem services in the National Adaptation Programmes of Action. Climate Policy, 12(4), 393-409. Prowse, M., N. Grist, and C. Sourang, 2009: Closing the gap between climate adaptation and poverty reduction frameworks. ODI Policy Briefing No. 21, June 2009. Overseas Development Institute (ODI), London, UK, pp. 4. Purdon, M., 2010: The clean development mechanism and community forests in Sub-Saharan Africa: reconsidering Kyoto's "moral position" on biocarbon sinks in the carbon market. Environment, Development and Sustainability, 12(6), 1025-1050. Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Quinn, C.H., G. Ziervogel, A. Taylor, T. Takama, and F. Thomalla, 2011: Coping with multiple stresses in rural South Africa. Ecology and Society, 16(3), 2. Raebild, A., U.B. Hansen, and S. Kambou, 2012: Regeneration of Vitellaria paradoxa and Parkia biglobosa in a parkland in Southern Burkina Faso. Agroforestry Systems, 85(3), 443-453. Rahmstorf, S., 2007: A semi-empirical approach to projecting future sea-level rise. Science, 315(5810), 368-370. Rain, D., R. Engstrom, C. Ludlow, and S. Antos, 2011: Accra Ghana: A city vulnerable to flooding and drought- induced migration. Case study prepared for Cities and Climate Change: Global Report on Human Settlements 2011, pp. 21. Raleigh, C., 2010: Political Marginalization, Climate Change, and Conflict in African Sahel States. International Studies Review, 12(1), 69-86. Raleigh, C., 2011: The search for safety: The effects of conflict, poverty and ecological influences on migration in the developing world. Global Environmental Change, 21(SUPPL. 1), S82-S93. Raleigh, C. and D. Kniveton, 2012: Come rain or shine: An analysis of conflict and climate variability in East Africa. Journal of Peace Research, 49(1), 51-64. Ramirez, J., A. Jarvis, I. Van den Bergh, C. Staver, and D. Turner, 2011: Changing Climates: Effects on Growing Conditions for Banana and Plantain (Musa spp.) and Possible Responses. In: Crop Adaptation to Climate Change. [Yadav, S.S., R. Redden, J.L. Hatfield, H. Lotze-Campen, and A.J.W. Hall (eds.)]. Wiley-Blackwell, Oxford, UK, pp. 426-438. Raworth, K., 2008: Gender and climate adaptation. Briefing Paper. Oxfam UK, pp. 5. Raxworthy, C.J., R.G. Pearson, N. Rabibisoa, A.M. Rakotondrazafy, J.-B. Ramanamanjato, A.P. Raselimanana, S. Wu, R.A. Nussbaum, and D.A. Stone, 2008: Extinction vulnerability of tropical montane endemism from warming and upslope displacement: A preliminary appraisal for the highest massif in Madagascar. Global Change Biology, 14(8), 1703-1720. Reid, H., S. Huq, and L. Murray, 2010: Community Champions: Adapting to Climate Challenges. International Institute for Environment and Development (IIED), London, UK, pp. 106. Reid, H., L. Sahlén, J. Stage, and J. MacGregor, 2008: Climate change impacts on Namibia's natural resources and economy. Climate Policy, 8(5), 452-466. Reij, C., G. Tappan, and M. Smale, 2009: Agroenvironmental transformation in the Sahel: another kind of Green Revolution . IFPRI Discussion Paper 00914, November 2009. International Food Policy Research Institute, Washington, DC, USA., pp. 43. Reiter, P., 2008: Global warming and malaria: Knowing the horse before hitching the cart. Malaria Journal, 7(SUPPL. 1), S3. Repetto, R., 2008: The climate crisis and the adaptation myth. Working Paper Number 13. Yale School of Forestry and Environmental Studies, New Haven, Conneticut, USA, pp. 20. Research Institute for Groundwater, 2011: Adaptation to the impact of sea level Rise in the Nile Delta Coastal zone, Egypt. Second Progress Report. Research Institute for Groundwater, National Water Research Center, Egypt. Reuveny, R., 2007: Climate change-induced migration and violent conflict. Political Geography, 26(6), 656-673. Reyburn, R., D.R. Kim, M. Emch, A. Khatib, L. Von Seidlein, and M. Ali, 2011: Climate variability and the outbreaks of cholera in Zanzibar, East Africa: A time series analysis. American Journal of Tropical Medicine and Hygiene, 84(6), 862-869. Rhajaoui, M., 2011: Human leishmaniases in Morocco: A nosogeographical diversity. Pathologie Biologie, 59(4), 226-229. Ricci, L., 2012: Peri-Urban Livelihood and Adaptive Capacity: Urban Development in Dar Es Salaam. Consilience: The Journal of Sustainable Development, 7(1), 46-63. Richard, K., D.M. Mwayafu, and H. Smith, 2011: REDD+ and adaptation to climate change in East Africa. Networking for equity in forest climate policy. REDD-net programme, pp. 8. Roberts, D., 2008: Thinking globally, acting locally - Institutionalizing climate change at the local government level in Durban, South Africa. Environment and Urbanization, 20(2), 521-537. Roberts, D., R. Boon, N. Diederichs, E. Douwes, N. Govender, A. Mcinnes, C. Mclean, S. O'Donoghue, and M. Spires, 2012: Exploring ecosystem-based adaptation in Durban, South Africa: "learning-by-doing" at the local government coal face. Environment and Urbanization, 24(1), 167-195. Roberts, L., 2010: The beginning of the end for Africa's devastating meningitis outbreaks? Science, 330(6010), 1466-1467. Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Robledo, C., N. Clot, A. Hammill, and B. Riché, 2012: The role of forest ecosystems in community-based coping strategies to climate hazards: Three examples from rural areas in Africa. Forest Policy and Economics, 24, 20- 28. Rocha, A., P. Melo-Gonçalves, C. Marques, J. Ferreira, and J.M. Castanheira, 2008: High-frequency precipitation changes in southeastern Africa due to anthropogenic forcing. International Journal of Climatology, 28(9), 1239- 1253. Rodima-Taylor, D., 2012: Social innovation and climate adaptation: Local collective action in diversifying Tanzania. Applied Geography, 33(1), 128-134. Rodó, X., M. Pascual, G. Fuchs, and A.S.G. Faruque, 2002: ENSO and cholera: A nonstationary link related to climate change? Proceedings of the National Academy of Sciences of the United States of America, 99(20), 12901-12906. Roehrig, R., D. Bouniol, F. Guichard, F. Hourdin, and J.C. Redelsperger, 2013: The present and future of the west african monsoon: A process-oriented assessment of CMIP5 simulations along the AMMA transect. Journal of Climate, 26(17), 6471-6505. Rogers, D.J. and S.E. Randolph, 2006: Climate Change and Vector-Borne Diseases , pp. 345-381. Romero González, A.M., A. Belemvire, and S. Sauliere, 2011: Climate Change and Women Farmers in Burkina Faso: Impact and adaptation policies and practices. Oxfam Research Reports. Oxfam International, Oxford, UK, pp. 45. Roncoli, C., B. Okoba, V. Gathaara, J. Ngugi, and T. Nganga, 2010: Adaptation to climate change for smallholder agriculture in Kenya: community-based perspectives from five districts. IFPRI Note. International Food Policy Research Institute (IFPRI), Washington, DC, USA, pp. 4. Roncoli, C., K. Ingram, and P. Kirshen, 2002: Reading the rains: Local knowledge and rainfall forecasting in Burkina Faso. Society and Natural Resources, 15(5), 409-427. Rosell, S. and B. Holmer, 2007: Rainfall change and its implications for Belg harvest in South Wollo, Ethiopia. Geografiska Annaler, Series A: Physical Geography, 89 A(4), 287-299. Rosenthal, D.M. and D.R. Ort, 2012: Examining Cassava's Potential to Enhance Food Security Under Climate Change. Tropical Plant Biology, 5(1), 30-38. Rosenzweig, C., G. Casassa, D.J. Karoly, A. Imeson, C. Liu, A. Menzel, S. Rawlins, T.L. Root, B. Seguin, and P. Tryjanowski, 2007: Assessment of observed changes and responses in natural and managed systems. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 79- 131. Roudier, P., B. Sultan, P. Quirion, and A. Berg, 2011: The impact of future climate change on West African crop yields: What does the recent literature say? Global Environmental Change, 21(3), 1073-1083. Rowell, D.P., 2012: Sources of uncertainty in future changes in local precipitation. Climate Dynamics, 39(7-8), 1929-1950. Rowhani, P., O. Degomme, D. Guha-Sapir, and E.F. Lambin, 2011: Malnutrition and conflict in East Africa: The impacts of resource variability on human security. Climatic Change, 105(1), 207-222. Ruelland, D., A. Tribotte, C. Puech, and C. Dieulin, 2011: Comparison of methods for LUCC monitoring over 50 years from aerial photographs and satellite images in a Sahelian catchment. International Journal of Remote Sensing, 32(6), 1747-1777. Ruelland, D., S. Ardoin-Bardin, L. Collet, and P. Roucou, 2012: Simulating future trends in hydrological regime of a large Sudano-Sahelian catchment under climate change. Journal of Hydrology, 424-425, 207-216. Ruppel, O.C., 2009: Regional economic communities and human rights in East and southern Africa. In: Human Rights in Africa. Legal Perspectives on their Protection and Promotion. [Bösl, A. and J. Diescho (eds.)]. Macmillan Education Namibia, Windhoek, Namibia, pp. 275-317. Ruppel, O.C., 2012: Climate change policy positions and related developments in the AU and SADC. SADC Law Jounal, 2(1), 14-35. Ruppel, O.C. and K. Ruppel-Schlichting, 2012: Climate change and human security: relevant for regional integration in SADC? In: Monitoring Regional Integration in Southern Africa 2011: Yearbook for Regional Integration. [Hartzenberg, T., G. Erasmus, and A. Du Pisani (eds.)]. Trade Law Centre for Southern Africa, Stellenbosch, South Africa, pp. 32-71. Subject to Final Copyedit 89 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Saarinen, J., W.L. Hambira, J. Atlhopheng, and H. Manwa, 2012: Tourism industry reaction to climate change in Kgalagadi South District, Botswana. Development Southern Africa, 29(2), 273-285. Saeed, F., A. Haensler, T. Weber, S. Hagemann, and D. Jacob, 2013: Representation of Extreme Precipitation Events Leading to Opposite Climate Change Signals over the Congo Basin. Atmosphere, 4, 254-271. Salah, A.B., Y. Kamarianakis, S. Chlif, N.B. Alaya, and P. Prastacos, 2007: Zoonotic cutaneous leishmaniasis in central Tunisia: Spatio-temporal dynamics. International Journal of Epidemiology, 36(5), 991-1000. Sallu, S.M., C. Twyman, and L.C. Stringer, 2010: Resilient or vulnerable livelihoods? assessing livelihood dynamics and trajectories in rural Botswana. Ecology and Society, 15(4), 3. Sanchez, A.C., P.E. Osborne, and N. Haq, 2011: Climate change and the African baobab (Adansonia digitata L.): The need for better conservation strategies. African Journal of Ecology, 49(2), 234-245. Sanderson, M.G., D.L. Hemming, and R.A. Betts, 2011: Regional temperature and precipitation changes under warming. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1934), 85-98. Sarmento, H., M. Isumbisho, and J.-P. Descy, 2006: Phytoplankton ecology of Lake Kivu (eastern Africa). Journal of Plankton Research, 28(9), 815-829. Sarvala, J., V.T. Langenberg, K. Salonen, D. Chitamwebwa, G.W. Coulter, T. Huttula, P. Kotilainen, N. Mulimbwa, and H. Mölsä, 2006: Changes in dissolved silica and transparency are not sufficient evidence for decreased primary productivity due to climate warming in Lake Tanganyika. Reply to comment by Verburg, Hecky and Kling. Verhandlungen Der Internationalen Vereiningung Für Theoretische Und Angewandte Limnologie, 29, 2339-2342. Satterthwaite, D., S. Huq, H. Reid, P. Pelling, and P. Romero Lankao, 2009: Adapting to climate change in urban areas: the possibilities and constraints in low- and middle-income nations. In: Adapting Cities to Climate Change. Understanding and Addressing the Development Challenges. [Bicknell, J., D. Dodman, and D. Satterthwaite (eds.)]. Routledge, London, UK, pp. 3-50. Saugeon, C., T. Baldet, M. Akogbeto, and M.C. Henry, 2009: Will climate and demography have a major impact on malaria in sub-Saharan Africa in the next 20 years? Médecine Tropicale : Revue Du Corps De Santé Colonial, 69(2), 203-207. Scheffran, J., E. Marmer, and P. Sow, 2012: Migration as a contribution to resilience and innovation in climate adaptation: Social networks and co-development in Northwest Africa. Applied Geography, 33(1), 119-127. Scheiter, S. and S.I. Higgins, 2009: Impacts of climate change on the vegetation of Africa: An adaptive dynamic vegetation modelling approach. Global Change Biology, 15(9), 2224-2246. Schilling, J., K.P. Freier, E. Hertig, and J. Scheffran, 2012: Climate change, vulnerability and adaptation in North Africa with focus on Morocco. Agriculture, Ecosystems and Environment, 156, 12-26. Schlenker, W. and D.B. Lobell, 2010: Robust negative impacts of climate change on African agriculture. Environmental Research Letters, 5(1). Schoneveld, G.C., L.A. German, and E. Nutako, 2011: Land-based investments for rural development? A grounded analysis of the local impacts of biofuel feedstock plantations in Ghana. Ecology and Society, 16(4), 10. Schreck, C.J. and F.H.M. Semazzi, 2004: Variability of the recent climate of eastern Africa. International Journal of Climatology, 24(6), 681-701. Scott, D., K. Senker, and E.C. England, 1982: Epidemiology of human Schistosoma haematobium infection around Volta Lake, Ghana, 1973-75. Bulletin of the World Health Organization, 60(1), 89-100. Seballos, F., T. Tanner, M. Tarazona, and J. Gallegos, 2011: Children and Disasters: Understanding Impact and Enabling Agency. Research Report. Children in a Changing Climate, Institute of Development Studies (IDS), Brighton, UK, pp. 59. SEI, 2009: The economics of climate change in Kenya. Final Report submitted in advance of COP15. Stockholm Environment Institute, Stockholm, Sweden, pp. 82. Sendzimir, J., C.P. Reij, and P. Magnuszewski, 2011: Rebuilding Resilience in the Sahel: Regreening in the Maradi and Zinder Regions of Niger. Ecology and Society, 16(3), 1. Seneviratne, S.I., N. Nicholls, D. Easterling, C.M. Goodess, S. Kanae, J. Kossin, Y. Luo, J. Marengo, K. McInnes, M. Rahimi, M. Reichstein, A. Sorteberg, C. Vera, and X. Zhang, 2012: Changes in climate extremes and their impacts on the natural physical environment. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX). Special Report of the Intergovernmental Panel on Climate Change (IPCC). [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al. (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 109-230. Subject to Final Copyedit 90 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Seo, S.N. and R. Mendelsohn, 2008: Measuring impacts and adaptations to climate change: A structural Ricardian model of African livestock management. Agricultural Economics, 38(2), 151-165. Seth, A., S.A. Rauscher, M. Rojas, A. Giannini, and S.J. Camargo, 2011: Enhanced spring convective barrier for monsoons in a warmer world? Climatic Change, 104(2), 403-414. Seto, K.C., 2011: Exploring the dynamics of migration to mega-delta cities in Asia and Africa: Contemporary drivers and future scenarios. Global Environmental Change, 21(SUPPL. 1), S94-S107. Shackleton, C.M., S.E. Shackleton, J. Gambiza, E. Nel, K. Rowntree, P. Urquhart, C. Fabricius, and A. Ainslie, 2010: Livelihoods and Vulnerability in the Arid and Semi-Arid Lands of Southern Africa: Exploring the Links Between Ecosystem Services & Poverty Alleviation. Nova Science Publishers, New York, pp. 267. Shackleton, S.E. and C.M. Shackleton, 2012: Linking poverty, HIV/AIDS and climate change to human and ecosystem vulnerability in southern Africa: consequences for livelihoods and sustainable ecosystem management. International Journal of Sustainable Development & World Ecology, 19(3), 275-286. Shankland, A. and R. Chambote, 2011: Prioritising PPCR Investments in Mozambique: The Politics of 'Country Ownership' and 'Stakeholder Participation'. IDS Bulletin, 42(3), 62-69. Shephard, G.S., 2008: Impact of mycotoxins on human health in developing countries. Food Additives and Contaminants, 25(2), 146-151. Shongwe, M.E., G.J. Van Oldenborgh, B.J.J.M. Van Den Hurk, B. De Boer, C.A.S. Coelho, and M.K. Van Aalst, 2009: Projected changes in mean and extreme precipitation in Africa under global warming. Part I: Southern Africa. Journal of Climate, 22(13), 3819-3837. Shongwe, M.E., G.J. van Oldenborgh, B. van den Hurk, and M. van Aalst, 2011: Projected changes in mean and extreme precipitation in Africa under global warming. Part II: East Africa. Journal of Climate, 24(14), 3718- 3733. Siegel, P.B., 2011: A 'no-regrets' risk-based approach to climate proofing public infrastructure: improved national and subnational planning for resilience and sustainable growth. In: Paving the way for climate-resilient infrastructure: Guidance for practitioners and planners. United Nations Development Programme (UNDP), New York, NY, USA, pp. 17-51. Sietz, D., M. Boschütz, and R.J.T. Klein, 2011: Mainstreaming climate adaptation into development assistance: Rationale, institutional barriers and opportunities in Mozambique. Environmental Science and Policy, 14(4), 493-502. Sillmann, J. and E. Roeckner, 2008: Indices for extreme events in projections of anthropogenic climate change. Climatic Change, 86(1-2), 83-104. Simane, B., B.F. Zaitchik, and D. Mesfin, 2012: Building climate resilience in the Blue Nile/Abay Highlands: A framework for action. International Journal of Environmental Research and Public Health, 9(2), 610-631. Sissoko, K., H. van Keulen, J. Verhagen, V. Tekken, and A. Battaglini, 2011: Agriculture, livelihoods and climate change in the West African Sahel. Regional Environmental Change, 11(SUPPL. 1), 119-125. Sitch, S., C. Huntingford, N. Gedney, P.E. Levy, M. Lomas, S.L. Piao, R. Betts, P. Ciais, P. Cox, P. Friedlingstein, C.D. Jones, I.C. Prentice, and F.I. Woodward, 2008: Evaluation of the terrestrial carbon cycle, future plant geography and climate-carbon cycle feedbacks using five Dynamic Global Vegetation Models (DGVMs). Global Change Biology, 14(9), 2015-2039. Smith, J., L. Deck, B. McCarl, P. Kirshen, J. Malley, and M. Abdrabo, 2013: Potential Impacts of Climate Change on the Egyptian Economy. United Nations Development Programme (UNDP), Cairo, Egypt, pp. 142. Smith, J.B., T. Dickinson, J.D.B. Donahue, I. Burton, E. Haites, R.J.T. Klein, and A. Patwardhan, 2011: Development and climate change adaptation funding: coordination and integration. Climate Policy, 11(3), 987- 1000. Smucker, T.A. and B. Wisner, 2008: Changing household responses to drought in Tharaka, Kenya: Vulnerability, persistence and challenge. Disasters, 32(2), 190-215. Snoussi, M., T. Ouchani, A. Khouakhi, and I. Niang-Diop, 2009: Impacts of sea-level rise on the Moroccan coastal zone: Quantifying coastal erosion and flooding in the Tangier Bay. Geomorphology, 107(1-2), 32-40. Solomon, T.B., H.A. Snyman, and G.N. Smit, 2007: Cattle-rangeland management practices and perceptions of pastoralists towards rangeland degradation in the Borana zone of southern Ethiopia. Journal of Environmental Management, 82(4), 481-494. Somorin, O.A., H.C.P. Brown, I. Visseren-Hamakers, D.J. Sonwa, B. Arts, and J. Nkem, 2012: The congo basin forests in a changing climate: Policy discourses on adaptation and mitigation (REDD+). Global Environmental Change, 22(1), 288-298. Subject to Final Copyedit 91 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sonwa, D., Y. Bele, O. O. Somorin, C. Jum, and J. Nkem, 2009: Adaptation for forests and communities in the Congo Basin. ETFRN News, 50, 93-100. Sonwa, D.J., J.N. Nkem, M.E. Idinoba, M.Y. Bele, and C. Jum, 2012: Building Regional Priorities in Forests for Development and Adaptation to Climate Change in the Congo Basin. Mitigation and Adaptation Strategies for Global Change, 17(4), 441-450. Sovacool, B.K., 2011: Hard and soft paths for climate change adaptation. Climate Policy, 11(4), 1177-1183. START, 2011: Enhancing scientific capacity to inspire informed action on global environmental change. START Annual Report 2010-2011. START International, Washington, DC, USA, pp. 36. Staver, A.C., S. Archibald, and S. Levin, 2011a: Tree cover in sub-Saharan Africa: Rainfall and fire constrain forest and savanna as alternative stable states. Ecology, 92(5), 1063-1072. Staver, A.C., S. Archibald, and S.A. Levin, 2011b: The global extent and determinants of savanna and forest as alternative biome states. Science, 334(6053), 230-232. Stensgaard, A.-S., J. Utzinger, P. Vounatsou, E. Hürlimann, N. Schur, C.F.L. Saarnak, C. Simoonga, P. Mubita, N.B. Kabatereine, L.-A.T. Tchuenté, C. Rahbek, and T.K. Kristensen, 2011: Large-scale determinants of intestinal schistosomiasis and intermediate host snail distribution across Africa: Does climate matter? Acta Tropica, in press. Stenuite, S., S. Pirlot, M.-A. Hardy, H. Sarmento, A.-L. Tarbe, B. Leporcq, and J.-P. Descy, 2007: Phytoplankton production and growth rate in Lake Tanganyika: Evidence of a decline in primary productivity in recent decades. Freshwater Biology, 52(11), 2226-2239. Stern, D.I., P.W. Gething, C.W. Kabaria, W.H. Temperley, A.M. Noor, E.A. Okiro, G.D. Shanks, R.W. Snow, and S.I. Hay, 2011: Temperature and malaria trends in highland East Africa. PLoS ONE, 6(9). Stevenson, D., R. Doherty, M. Sanderson, C. Johnson, B. Collins, and D. Derwent, 2005: Impacts of climate change and variability on tropospheric ozone and its precursors. Faraday Discussions, 130, 41-57. Steynor, A.C., B.C. Hewitson, and M.A. Tadross, 2009: Projected future runoff of the Breede river under climate change. Water SA, 35(4), 433-440. Stott, P.A., N.P. Gillett, G.C. Hegerl, D.J. Karoly, D.A. Stone, X. Zhang, and F. Zwiers, 2010: Detection and attribution of climate change: A regional perspective. Wiley Interdisciplinary Reviews: Climate Change, 1(2), 192-211. Stott, P.A., G.S. Jones, N. Christidis, F.W. Zwiers, G. Hegerl, and H. Shiogama, 2011: Single-step attribution of increasing frequencies of very warm regional temperatures to human influence. Atmospheric Science Letters, 12(2), 220-227. Stott, P.A., 2003: Attribution of regional-scale temperature changes to anthropogenic and natural causes. Geophysical Research Letters, 30(14), 1728. Stringer, L.C., J.C. Dyer, M.S. Reed, A.J. Dougill, C. Twyman, and D. Mkwambisi, 2009: Adaptations to climate change, drought and desertification: local insights to enhance policy in southern Africa. Environmental Science and Policy, 12(7), 748-765. Stryker, J.J. and A. Bomblies, 2012: The impacts of land use change on malaria vector abundance in a water-limited, highland region of Ethiopia. EcoHealth, 9(4), 455-470. Suarez, P., F. Ching, G. Ziervogel, I. Lemaire, D. Turnquest, J.M. de Suarez, and B. Wisner, 2008: Video-mediated approaches for community-level climate adaptation. IDS Bulletin, 39(4), 96-104. Sultan, B., K. Labadi, J.-F. Guégan, and S. Janicot, 2005: Climate drives the meningitis epidemics onset in West Africa. PLoS Medicine, 2(1), e6. Sultan, B., P. Roudier, P. Quirion, A. Alhassane, B. Muller, M. Dingkuhn, P. Ciais, M. Guimberteau, S. Traore, and C. Baron, 2013: Assessing climate change impacts on sorghum and millet yields in the Sudanian and Sahelian savannas of West Africa. Environmental Research Letters, 8(1). Syampungani, S., P.W. Chirwa, F.K. Akinnifesi, and O.C. Ajayi, 2010: The Potential of Using Agroforestry as a Win-Win Solution to Climate Change Mitigation and Adaptation and Meeting Food Security Challenges in Southern Africa. Agricultural Journal, 5(2), 80-88. Sylla, M.B., A.T. Gaye, and G.S. Jenkins, 2012: On the fine-scale topography regulating changes in atmospheric hydrological cycle and extreme rainfall over West Africa in a regional climate model projections. International Journal of Geophysics, 2012, Article no. 981649. Tacoli, C., 2011: Not only climate change: mobility, vulnerability and socio-economic transformations in environmentally fragile areas of Bolivia, Senegal and Tanzania. Human Settlements Working Paper No. 28: Subject to Final Copyedit 92 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rural-Urban Interactions and Livelihood Strategies. International Institute for Environment and Development (IIED), London, UK, pp. 39. Tacoli, C., 2009: Crisis or adaptation? Migration and climate change in a context of high mobility. Environment and Urbanization, 21(2), 513-525. Tadross, M., P. Suarez, A. Lotsch, S. Hachigonta, M. Mdoka, L. Unganai, F. Lucio, D. Kamdonyo, and M. Muchinda, 2009: Growing-season rainfall and scenarios of future change in southeast Africa: Implications for cultivating maize. Climate Research, 40(2-3), 147-161. Tadross, M.A., B.C. Hewitson, and M.T. Usman, 2005: The interannual variability on the onset of the maize growing season over South Africa and Zimbabwe. Journal of Climate, 18(16), 3356-3372. Tanner, T. and T. Mitchell, 2008: Entrenchment or enhancement: could climate change adaptation help reduce poverty?. Working Paper No. 106. Chronic Poverty Research Centre, Brighton, UK, pp. 18. Tanner, T. and F. Seballos, 2012: Children, Climate Change and Disasters. IDS In Focus Policy Briefing 23. Institute of Development Studies (IDS), Brighton, UK, pp. 4. Tawodzera, G., 2011: Vulnerability in crisis: Urban household food insecurity in Epworth, Harare, Zimbabwe. Food Security, 3(4), 503-520. Taye, M.T., V. Ntegeka, N.P. Ogiramoi, and P. Willems, 2011: Assessment of climate change impact on hydrological extremes in two source regions of the Nile River Basin. Hydrology and Earth System Sciences, 15(1), 209-222. Taylor, R.G., A.D. Koussis, and C. Tindimuguya, 2009: Groundwater and climate in Africa - A review. Hydrological Sciences Journal, 54(4), 655-664. Tebaldi, C., K. Hayhoe, J.M. Arblaster, and G.A. Meehl, 2006: Going to the extremes: An intercomparison of model-simulated historical and future changes in extreme events. Climatic Change, 79(3-4), 185-211. Tefera, T., F. Kanampiu, H. De Groote, J. Hellin, S. Mugo, S. Kimenju, Y. Beyene, P.M. Boddupalli, B. Shiferaw, and M. Banziger, 2011: The metal silo: An effective grain storage technology for reducing post-harvest insect and pathogen losses in maize while improving smallholder farmers' food security in developing countries. Crop Protection, 30(3), 240-245. Terblanche, J.S., S. Clusella-Trullas, J.A. Deere, and S.L. Chown, 2008: Thermal tolerance in a south-east African population of the tsetse fly Glossina pallidipes (Diptera, Glossinidae): Implications for forecasting climate change impacts. Journal of Insect Physiology, 54(1), 114-127. Theisen, O.M., 2012: Climate clashes? weather variability, land pressure, and organized violence in Kenya, 1989- 2004. Journal of Peace Research, 49(1), 81-96. Theron, A. and M. Rossouw, 2008: Analysis of potential coastal zone climate change impacts and possible response options in the southern African region. Proceedings of Science real and relevant: 2nd CSIR Biennial Conference, 17 & 18 November 2008, CSIR International Convention Centre Pretoria, pp. 10. Thierfelder, C. and P.C. Wall, 2010: Investigating conservation agriculture (CA) systems in Zambia and Zimbabwe to mitigate future effects of climate change. Journal of Crop Improvement, 24(2), 113-121. Thomas, D.S.G. and C. Twyman, 2005: Equity and justice in climate change adaptation amongst natural-resource- dependent societies. Global Environmental Change, 15(2), 115-124. Thomas, D.S.G., C. Twyman, H. Osbahr, and B. Hewitson, 2007: Adaptation to climate change and variability: Farmer responses to intra-seasonal precipitation trends in South Africa. Climatic Change, 83(3), 301-322. Thomas, R.J., 2008: Opportunities to reduce the vulnerability of dryland farmers in Central and West Asia and North Africa to climate change. Agriculture, Ecosystems and Environment, 126(1-2), 36-45. Thomson, A.M., K.V. Calvin, L.P. Chini, G. Hurtt, J.A. Edmonds, B. Bond-Lamberty, S. Frolking, M.A. Wise, and A.C. Janetos, 2010: Climate mitigation and the future of tropical landscapes. Proceedings of the National Academy of Sciences of the United States of America, 107(46), 19633-19638. Thomson, M.C., A.M. Molesworth, M.H. Djingarey, K.R. Yameogo, F. Belanger, and L.E. Cuevas, 2006: Potential of environmental models to predict meningitis epidemics in Africa. Tropical Medicine and International Health, 11(6), 781-788. Thorn, J., 2011: A participatory investigation into community-based climate adaptation to urban flooding: the case of Mathare Valley Slums, Nairobi. Diss. MSc, School of Geography and Environment, St Catherine College, University of Oxford, Oxford, UK, 49 pp. Thornton, P.K., P.G. Jones, G. Alagarswamy, and J. Andresen, 2009a: Spatial variation of crop yield response to climate change in East Africa. Global Environmental Change, 19(1), 54-65. Subject to Final Copyedit 93 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Thornton, P.K., J. van de Steeg, A. Notenbaert, and M. Herrero, 2009b: The impacts of climate change on livestock and livestock systems in developing countries: A review of what we know and what we need to know. Agricultural Systems, 101(3), 113-127. Thornton, P.K., P.G. Jones, G. Alagarswamy, J. Andresen, and M. Herrero, 2010: Adapting to climate change: Agricultural system and household impacts in East Africa. Agricultural Systems, 103(2), 73-82. Thornton, P.K., P.G. Jones, P.J. Ericksen, and A.J. Challinor, 2011: Agriculture and food systems in sub-Saharan Africa in a 4°C+ world. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1934), 117-136. Thurlow, J. and P. Wobst, 2003: Poverty Focused Social Accounting Matrices for Tanzania. TMD Discussion Paper No. 112. International Food Policy Research Institute (IFPRI) - Trade and Macroeconomics Division (TMD), Washington DC, pp. 59. Tierney, J.E., M.T. Mayes, N. Meyer, C. Johnson, P.W. Swarzenski, A.S. Cohen, and J.M. Russell, 2010: Late- twentieth-century warming in Lake Tanganyika unprecedented since AD 500. Nature Geoscience, 3(6), 422- 425. Tingem, M. and M. Rivington, 2009: Adaptation for crop agriculture to climate change in Cameroon: Turning on the heat. Mitigation and Adaptation Strategies for Global Change, 14(2), 153-168. Tompkins, E.L. and W.N. Adger, 2003: Building resilience to climate change through adaptive management of natural resources. Tyndall Centre Working Paper No. 27. Tyndall Centre for Climate Change Research, Norwich, UK, pp. 19. Tonnang, H.E.Z., R.Y.M. Kangalawe, and P.Z. Yanda, 2010: Predicting and mapping malaria under climate change scenarios: The potential redistribution of malaria vectors in Africa. Malaria Journal, 9, 111. Toteng, E.N., 2012: Political and State System Barriers to the Implementation of Climate Change Adaptation in Botswana: Implications for SADC Regional Climate Change Policy . In: Overcoming Barriers to Climate Change Adaptation Implementation in Southern Africa. [Masters, L. and L. Duff (eds.)]. Africa Institute of South Africa, Pretoria, South Africa, pp. 22-28. Touchan, R., K.J. Anchukaitis, D.M. Meko, S. Attalah, C. Baisan, and A. Aloui, 2008: Long term context for recent drought in northwestern Africa. Geophysical Research Letters, 35(13). Tougiani, A., C. Guero, and T. Rinaudo, 2009: Community mobilisation for improved livelihoods through tree crop management in Niger. GeoJournal, 74(5), 377-389. Toumi, A., S. Chlif, J. Bettaieb, N.B. Alaya, A. Boukthir, Z.E. Ahmadi, and A.B. Salah, 2012: Temporal dynamics and impact of climate factors on the incidence of Zoonotic Cutaneous Leishmaniasis in central Tunisia. PLoS Neglected Tropical Diseases, 6(5). Trenberth, K.E., P.D. Jones, P. Ambenje, R. Bojariu, D. Easterling, A. Klein Tank, D. Parker, F. Rahimzadeh, J.A. Renwick, M. Rusticucci, B. Soden, and P. Zhai, 2007: Observations: Surface and Atmospheric Climate Change. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 236-336. Tschakert, P. and K.A. Dietrich, 2010: Anticipatory learning for climate change adaptation and resilience. Ecology and Society, 15(2), 11. Tschakert, P. and R. Tutu, 2010: Environment, Forced Migration and Social Vulnerability. In: Solastalgia: Environmentally Induced Distress and Migration Among Africa s Poor Due to Climate Change [Afifi, T. and J. Jäger(eds.)]. Springer, Berlin, Heidelberg, pp. 57-69. Tshimanga, R.M. and D.A. Hughes, 2012: Climate change and impacts on the hydrology of the Congo Basin: The case of the northern sub-basins of the Oubangui and Sangha Rivers. Physics and Chemistry of the Earth, 50-52, 72-83. Tumbare, M.J., 2008: Managing Lake Kariba sustainably: Threats and challenges. Management of Environmental Quality, 19(6), 731-739. UN, AfDB, AU, EC, IMF, IDB, OECD, and The World Bank, 2008: Achieving the Millennium Development Goals in Africa. Recommendations of the MDG Africa Steering Group. United Nations, New York, NY, USA, pp. 39. UN, 2011: The Millennium Development Goals Report. United Nations, New York, NY, USA, pp. 67. UN DESA, 2010: World Urbanization Prospects: The 2009 Revision. United Nations Department of Economic and Social Affairs, Population Division, New York, pp. 47. Subject to Final Copyedit 94 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 UN DESA, 2011: World Population Prospects: The 2010 Revision. Highlights and Advance Tables. United Nations Department of Economic and Social Affairs, Population Division, New York, NY, USA, pp. 142. UN DESA, 2013: World Population Prospects: The 2012 Revision. Excel Tables - Population Data. Accessed at: http://esa.un.org/unpd/wpp/Excel-Data/population.htm. United Nations Department of Economic and Social Affairs (UN DESA), Population Division. UNCCD, UNDP, and UNEP, 2009: Climate change in the African drylands: options and opportunities for adaptation and mitigation. United Nations Convention to Combat Desertification (UNCCD), United Nations Development Programme (UNDP) and United Nations Environment Programme (UNEP), Bonn, Germany / New York, USA / Nairobi, Kenya, pp. 54. UNDP, 2009: Rapport mondial sur le développement humain 2009. Levers les barrieres: Mobilité et développement humains. Programme des Nations Unies pour le Développement (PNUD), New York, NY, USA, pp. 237. UNDP, 2011a: Down to earth: territorial approach to climate change. Low-emission and climate-resilient development strategies at the sub-national level. 2011 Update. United Nations Development Programme (UNDP), New York, NY, pp. 18. UNDP, 2011b: Africa Adaptation Programme experiences. Gender and climate change: advancing development through an integrated gender perspective. Discussion Paper Series, Vol. 1, March 2011. United Nations Development Programme (UNDP), New York, NY, USA, pp. 16. UNDP, ECA, AfDB, and AUC, 2011: Assessing progress in Africa toward the Millennium Development Goals: MDG Report 2011. United Nations Economic Commission for Africa (ECA); African Union Commission (AUC); African Development Bank (AfDB); United Nation Development Programme Regional Bureau for Africa (UNDP RBA), pp. 136. UNDP-UNEP Poverty-Environment Initiative, 2011a: Enabling local success: a primer on mainstreaming local ecosystem-based solutions to poverty-environment challenges. UNDP-UNEP Poverty-Environment Facility, Nairobi, Kenya, pp. 65. UNDP-UNEP Poverty-Environment Initiative, 2011b: Mainstreaming climate change adaptation into development planning: a guide for practitioners. UNDP-UNEP Poverty-Environment Initiative, Nairobi, Kenya, pp. 86. UNEP and IOC-UNESCO, 2009: Summary for decision makers. In: An Assessment of Assessments. Findings of the Group of Experts. Start-up Phase of a Regular Process for Global Reporting and Assessment of the State of the Marine Environment including Socio-economic Aspects. United Nations Environment Programme (UNEP) and Intergovernmental Oceanographic Commission of UNESCO (IOC-UNESCO), Nairobi, Kenya, pp. 44. UNFCCC, 2007: Report on the African regional workshop on adaptation: Note by the secretariat. FCCC/SBI/2007/2, 12 February 2007. United Nations Framework Convention on Climate Change (UNFCCC), pp. 15. UNFCCC, 2009: Copenhagen Accord . UNFCCC, 2011: The Cancun Agreements: Outcome of the work of the Ad Hoc Working Group on Long-term Cooperative Action under the Convention . UN-Habitat, 2008: State of the world's cities 2008/2009. Harmonious cities. Earthscan, London, UK and Sterling, VA, USA, pp. 259. UN-Habitat and UNEP, 2010: The state of African cities 2010: governance, inequality and urban land markets. UN- Habitat and United Nations Environment Programme (UNEP), Nairobi, Kenya, pp. 268. UN-Habitat, 2011: Cities and Climate Change: Global Report on Human Settlements 2011. United Nations Human Settlements Programme (UN-Habitat) and Earthscan, London, UK and Nairobi, Kenya, pp. 279. UNICEF, 2007: Climate change and children. United Nations Children s Fund (UNICEF), New York, NY, USA, pp. 20. UNICEF and WHO, 2008: Progress on Drinking Water and Sanitation: Special Focus on Sanitation. Joint Monitoring Programme for Water Supply and Sanitation (JMP). United Nations Children s Fund (UNICEF) and World Health Organization (WHO), New York, NY, USA and Geneva, Switzerland, pp. 54. UNICEF, 2010: Climate change in Kenya: Focus on children. United Nations Children's Fund (UNICEF) UK, London, UK, pp. 28. UNICEF, 2011: Exploring the impact of climate change on children in South Africa. United Nations Children's Fund (UNICEF) South Africa, Pretoria, South Africa, pp. 119. UNICEF and WHO, 2012: Progress on Drinking Water and Sanitation: 2012 Update. WHO/UNICEF Joint Monitoring Programme for Water Supply and Sanitation. United Nations Children's Fund (UNICEF) and the World Health Organization (WHO), New York, NY, USA, pp. 59. Subject to Final Copyedit 95 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 UNISDR, 2011: Effective measures to build resilience in Africa to adapt to climate change. Briefing Note 04. United Nations International Strategy for Disaster Reduction (UNISDR), Geneva, Switzerland, pp. 8. UN-OCAH and IDMC, 2009: Monitoring disaster displacement in the context of climate change. United Nations Office for the Coordination of the Humanitarian Affairs (UN-OCHA) and Internal Displacement Monitoring Centre (IDMC), Geneva, Switzerland, pp. 30. UNPFII, 2008: Meeting Report. Conference on indigenous peoples and climate change, Copenhagen, 21-22 February 2008. /C.19/2008/CRP. 3. United Nations Permanent Forum on Indigenous Issues (UNPFII), pp. 11. Unruh, J.D., 2008: Carbon sequestration in Africa: The land tenure problem. Global Environmental Change, 18(4), 700-707. Urdal, H., 2005: People vs. Malthus: Population pressure, environmental degradation, and armed conflict revisited. Journal of Peace Research, 42(4), 417-434. Urquhart, P., 2014: Governing the Ungovernable? Climate Change and Social Justice in Southern Africa. In: Governance for Justice and Environmental Sustainability: Lessons across Natural Resource Sectors in Sub- Saharan Africa. [Sowman, M. and R. Wynberg (eds.)]. Routledge. Urquhart, P., 2009: IFAD s response to climate change through support to adaptation and related actions. Comprehensive report: Final version. International Fund for Agricultural Development (IFAD), Rome, pp. 160. van Aalst, M.K., T. Cannon, and I. Burton, 2008: Community level adaptation to climate change: The potential role of participatory community risk assessment. Global Environmental Change, 18(1), 165-179. van Breugel, P., M. Herrero, J. van de Steeg, and D. Peden, 2010: Livestock water use and productivity in the Nile basin. Ecosystems, 13(2), 205-221. van de Giesen, N., J. Liebe, and G. Jung, 2010: Adapting to climate change in the Volta Basin, West Africa. Current Science, 98(8), 1033-1037. van der Geest, K. and T. Dietz, 2004: A literature survey about risk and vulnerability in drylands, with a focus on the Sahel. In: The impact of climate change on drylands: with a focus on West Africa. [Dietz, A.J., R. Rueben, and J. Verhagen (eds.)]. Kluwer Academic Publishers, Dordrecht, The Netherlands, pp. 117-146. Van der Geest, K., 2011: North-South Migration in Ghana: What Role for the Environment? International Migration, 49(SUPPL.1), e69-e94. van Vliet, N., 2010: Participatory vulnerability assessment in the context of conservation and development projects: A case study of local communities in Southwest Cameroon. Ecology and Society, 15(2), 6. Vandeweerd, V., Y. Glemarec, and S. Billett, 2012: Readiness for Climate Finance. A framework for understanding what it means to be ready to use climate finance. United Nations Development Programme (UNDP), New York, NY, USA, pp. 27. Vang Rasmussen, L., K. Rasmussen, T. Birch-Thomsen, S.B.P. Kristensen, and O. Traoré, 2012a: The effect of cassava-based bioethanol production on above-ground carbon stocks: A case study from Southern Mali. Energy Policy, 41, 575-583. Vang Rasmussen, L., K. Rasmussen, and T. Bech Bruun, 2012b: Impacts of Jatropha-based biodiesel production on above and below-ground carbon stocks: A case study from Mozambique. Energy Policy, 51, 728-736. Verburg, P., R.E. Hecky, and H.J. Kling, 2007: Climate warming decreased primary productivity in Lake Tanganyika, inferred from accumulation of dissolved silica and increased transparency. Comment to Sarvala et al. 2006. Verhandlungen Der Internationalen Vereiningung Für Theoretische Und Angewandte Limnologie, 29, 2335-2338. Verburg, P. and R.E. Hecky, 2009: The physics of the warming of Lake Tanganyika by climate change. Limnology and Oceanography, 54(6 PART 2), 2418-2430. Verchot, L.V., M. Van Noordwijk, S. Kandji, T. Tomich, C. Ong, A. Albrecht, J. Mackensen, C. Bantilan, K.V. Anupama, and C. Palm, 2007: Climate change: Linking adaptation and mitigation through agroforestry. Mitigation and Adaptation Strategies for Global Change, 12(5), 901-918. Vermuelen, S., K. Dossou, D. Macqueen, D. Walubengo, and E. Nangoma, 2008: Spring back: climate resilience at Africa s grassroots. Sustainable Development Opinion. International Institute for Environment and Development (IIED), London, UK, pp. 2. Viatte, G., J. De Graaf, M. Demeke, T. Takahatake, and M. Rey de Arce, 2009: Responding to the food crisis: synthesis of medium-term measures proposed in inter-agency assessments. Food and Agriculture Organization of the United Nations (FAO), Rome, pp. 100. Subject to Final Copyedit 96 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Vié, J., C. Hilton-Taylor, and S.N. Stuart (eds.), 2009: Wildlife in a changing world: An analysis of the 2008 IUCN Red List of Threatened Species International Union for Conservation of Nature (IUCN), Gland, Switzerland, pp. 157. Vincent, K., 2007: Uncertainty in adaptive capacity and the importance of scale. Global Environmental Change, 17(1), 12-24. Vincent, K., A. Joubert, T. Cull, J. Magrath, and P. Johnston, 2011a: Overcoming the barriers: How to ensure future food production under climate change in Southern Africa. Oxfam Research Report. Oxfam GB for Oxfam International, Oxford, UK, pp. 59. Vincent, L.A., E. Aguilar, M. Saindou, A.F. Hassane, G. Jumaux, D. Roy, P. Booneeady, R. Virasami, L.Y.A. Randriamarolaza, F.R. Faniriantsoa, V. Amelie, H. Seeward, and B. Montfraix, 2011b: Observed trends in indices of daily and extreme temperature and precipitation for the countries of the western Indian Ocean, 1961- 2008. Journal of Geophysical Research D: Atmospheres, 116(10). Vincke, C., I. Diédhiou, and M. Grouzis, 2010: Long term dynamics and structure of woody vegetation in the Ferlo (Senegal). Journal of Arid Environments, 74(2), 268-276. Vizy, E.K. and K.H. Cook, 2012: Mid-twenty-first-century changes in extreme events over northern and tropical Africa. Journal of Climate, 25(17), 5748-5767. Vizy, E.K., K.H. Cook, J. Crétat, and N. Neupane, 2013: Projections of a Wetter Sahel in the Twenty-First Century from Global and Regional Models. Journal of Climate, 26(13), 4664-4687. Vogel, C. and K. O'Brien, 2006: Who can eat information? Examining the effectiveness of seasonal climate forecasts and regional climate-risk management strategies. Climate Research, 33(1), 111-122. Vogel, C., S.C. Moser, R.E. Kasperson, and G.D. Dabelko, 2007: Linking vulnerability, adaptation, and resilience science to practice: Pathways, players, and partnerships. Global Environmental Change, 17(3-4), 349-364. Vohland, K. and B. Barry, 2009: A review of in situ rainwater harvesting (RWH) practices modifying landscape functions in African drylands. Agriculture, Ecosystems and Environment, 131(3-4), 119-127. Vörösmarty, C.J., E.M. Douglas, P.A. Green, and C. Revenga, 2005: Geospatial indicators of emerging water stress: An application to Africa. Ambio, 34(3), 230-236. Walker, N.J. and R.E. Schulze, 2008: Climate change impacts on agro-ecosystem sustainability across three climate regions in the maize belt of South Africa. Agriculture, Ecosystems and Environment, 124(1-2), 114-124. Warner, K., M. Hamza, A. Oliver-Smith, F. Renaud, and A. Julca, 2010: Climate change, environmental degradation and migration. Natural Hazards, 55(3), 689-715. Warner, K., K. van der Geest, S. Kreft, S. Huq, S. Harmeling, K. Kusters, and A. De Sherbinin, 2012: Evidence from the frontlines of climate change: Loss and damage to communities despite coping and adaptation. Policy Report No. 9. United Nations University Institute for Environment and Human Security (UNU-EHS), Bonn, Germany, pp. 85. Washington, R., M. Harrison, D. Conway, E. Black, A. Challinor, D. Grimes, R. Jones, A. Morse, G. Kay, and M. Todd, 2006: African climate change: Taking the shorter route. Bulletin of the American Meteorological Society, 87(10), 1355-1366. Watkiss, P., T.E. Downing, and J. Dyszynski, 2010: ADAPTCost Project: Analysis of the economic costs of climate change adaptation in Africa. United National Environment Programme (UNEP), Nairobi, Kenya, pp. 35. Watterson, I.G., 2009: Components of precipitation and temperature anomalies and change associated with modes of the Southern Hemisphere. International Journal of Climatology, 29(6), 809-826. Weiß, M., R. Schaldach, J. Alcamo, and M. Flörke, 2009: Quantifying the human appropriation of fresh water by African agriculture. Ecology and Society, 14(2), 25. Welcomme, R., 2011: Review of the state of world shery resources: Inland fisheries. FAO Fisheries and Aquaculture Circular No. 942, Rev. 2. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy, pp. 97. Wertz-Kanounnikoff, S., A. Sitoe, and A. Salomao, 2011: How is REDD+ unfolding in southern Africa's dry forests? A snapshot from Mozambique. CIFOR Infobrief No. 37. Center for International Forestry Research (CIFOR), Maputo, Mozambique, pp. 8. Westerhoff, L. and B. Smit, 2009: The rains are disappointing us: Dynamic vulnerability and adaptation to multiple stressors in the Afram Plains, Ghana. Mitigation and Adaptation Strategies for Global Change, 14(4), 317-337. Westgate, K., 2010: Toolkit for national platforms for disaster risk reduction in Africa: Contributing to the implementation of the Hyogo Framework for Action and the Africa Regional Strategy and Programme of Action Subject to Final Copyedit 97 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 for DRR. United Nations International Strategy for Disaster Reduction (UNISDR) - Regional Office for Africa, Nairobi, Kenya, pp. 48. WFP, 2011: Building resilience: Bridging food security, climate change adaptation and disaster risk reduction: An overview of workshop case studies. World Food Programme Office for Climate Change and Disaster Risk Reduction, Rome, pp. 35. WHO, 2007: Outbreaks of Rift Valley fever in Kenya, Somalia and United Republic of Tanzania, December 2006- April 2007. Relevé Épidémiologique Hebdomadaire / Section D'Hygiene Du Secrétariat De La Société Des Nations = Weekly Epidemiological Record / Health Section of the Secretariat of the League of Nations, 82(20), 169-178. WHO, 2009: Sommet africain sur les maladies tropicales négligées. Theme: Lutte contre la pauvreté et la maladie: donner de l espoir aux meres et aux enfants [French: combatting poverty and illness. Report of the African summit on neglected tropical diseases] 30 March -3 April 2009, Palais des Congres, Cotonou, Bénin, pp. 5. WHO, 2012: World Malaria Report 2012. World Health Organization (WHO), Geneva, Switzerland, pp. 195. WHO, 2013a: Global Health Observatory (GHO) - The data repository. Accessed at: http://www.who.int/gho/database/en/. World Health Organization (WHO), Geneva, Switzerland. WHO, 2013b: Schistosomiasis - Fact sheet N°115 World Health Organization (WHO), Geneva, Switzerland. Wigley, B.J., W.J. Bond, and M.T. Hoffman, 2009: Bush encroachment under three contrasting land-use practices in a mesic South African savanna. African Journal of Ecology, 47(SUPPL. 1), 62-70. Wigley, B.J., W.J. Bond, and M.T. Hoffman, 2010: Thicket expansion in a South African savanna under divergent land use: Local vs. global drivers? Global Change Biology, 16(3), 964-976. Williams, A.P. and C. Funk, 2011: A westward extension of the warm pool leads to a westward extension of the Walker circulation, drying eastern Africa. Climate Dynamics, 37(11-12), 2417-2435. Williams, A.P., C. Funk, J. Michaelsen, S.A. Rauscher, I. Robertson, T.H.G. Wils, M. Koprowski, Z. Eshetu, and N.J. Loader, 2012: Recent summer precipitation trends in the Greater Horn of Africa and the emerging role of Indian Ocean sea surface temperature. Climate Dynamics, 39(9-10), 2307-2328. Wiseman, W., J. Van Domelen, and S. Coll-Black, 2009: Designing and implementing a rural safety net in a low income setting: Lessons learned from Ethiopia s Productive Safety Net Program 2005-2009. The Productive Safety Net Program (PSNP), pp. 155. Wolski, P., M.C. Todd, M.A. Murray-Hudson, and M. Tadross, 2012: Multi-decadal oscillations in the hydro- climate of the Okavango River system during the past and under a changing climate. Journal of Hydrology, 475, 294-305. Woltering, L., A. Ibrahim, D. Pasternak, and J. Ndjeunga, 2011: The economics of low pressure drip irrigation and hand watering for vegetable production in the Sahel. Agricultural Water Management, 99(1), 67-73. Woodfine, A., 2009: Using sustainable land management practices to adapt to and mitigate climate change in sub- Saharan Africa: Resource Guide Version 1.0. TerrAfrica, pp. 79. World Bank, 2010: Report on the status of disaster risk reduction in sub-Saharan Africa. The World Bank, Africa Disaster Risk Management Team, Washington, DC, USA, pp. 44. World Bank and IMF, 2010: Global Monitoring Report 2010: The MDGs after the Crisis. World Bank and International Monetary Fund (IMF), pp. 151. World Bank, 2011: North African coastal cities address natural disasters and climate change. Climate Change Adaptation and Natural Disasters Preparedness in the Coastal Cities of North Africa, Summary of the Regional Study. The World Bank and Marseille Center for Mediterranean Integration (CMI), pp. 23. WTO Trade Policy Review, 2008: Trade policy review report by the Secretariat Ghana. WT/TPR/S/194. World Trade Organziation Trade Policy Review Body, pp. 10. Yabi, I. and F. Afouda, 2012: Extreme rainfall years in Benin (West Africa). Quaternary International, 262, 39-43. Yaka, P., B. Sultan, H. Broutin, S. Janicot, S. Philippon, and N. Fourquet, 2008: Relationships between climate and year-to-year variability in meningitis outbreaks: A case study in Burkina Faso and Niger. International Journal of Health Geographics, 7, 34. Yaro, J.A., 2010: The social dimensions of adaptation to climate change in Ghana. World Bank Discussion Paper No. 15, December 2010. The World Bank, Washinton, DC, USA, pp. 88. Yé, Y., V.R. Louis, S. Simboro, and R. Sauerborn, 2007: Effect of meteorological factors on clinical malaria risk among children: An assessment using village-based meteorological stations and community-based parasitological survey. BMC Public Health, 7, 101. Subject to Final Copyedit 98 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Yesuf, M., S. Di Falco, T. Deressa, C. Ringler, and G. Kohlin, 2008: The impact of climate change and adaptation on food production in low-income countries: evidence from the Nile Basin, Ethiopia. IFPRI Research Brief 15- 11. International Food Policy Research Institute (IFPRI), Washington, DC, USA, pp. 2. Yiougo, L., T. Sanon, and W.S. Ouedraogo, 2007: Facteurs de recrudescence de la leishmaniose cutanee et ses consequences socio economiques et sanitaires dans le secteur 30 de la ville de Ouagadougou (English: Factors affecting the resurgence of cutaneous leishmaniasis and its socioeconomic and health implications in Sector 30 of Ouagadougou city). Rapport de recherche. 2iE groupe EIER-ETSHER, pp. 103. You, L., C. Ringler, U. Wood-Sichra, R. Robertson, S. Wood, T. Zhu, G. Nelson, Z. Guo, and Y. Sun, 2011: What is the irrigation potential for Africa? A combined biophysical and socioeconomic approach. Food Policy, 36(6), 770-782. Yuen, B. and A. Kumssa (eds.), 2011: Climate Change and Sustainable Urban Development in Africa and Asia Springer, Dordrecht, Heidelberg, London, New York, pp. 278. Zelazowski, P., Y. Malhi, C. Huntingford, S. Sitch, and J.B. Fisher, 2011: Changes in the potential distribution of humid tropical forests on a warmer planet. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1934), 137-160. Zeng, G., J.A. Pyle, and P.J. Young, 2008: Impact of climate change on tropospheric ozone and its global budgets. Atmospheric Chemistry and Physics, 8(2), 369-387. Zhou, L., R.E. Dickinson, A. Dai, and P. Dirmeyer, 2010: Detection and attribution of anthropogenic forcing to diurnal temperature range changes from 1950 to 1999: Comparing multi-model simulations with observations. Climate Dynamics, 35(7), 1289-1307. Ziervogel, G. and A. Opere, 2010: Integrating meteorological and indigenous knowledge-based seasonal climate forecasts for the agricultural sector: Lessons from participatory action research in sub-Saharan Africa. Climate Change Adaptation in Africa Learning Paper Series. International Development Research Centre (IDRC), Ottawa, Canada, pp. 19. Ziervogel, G., A. Taylor, S. Hachigonta, and J. Hoffmaister, 2008: Climate adaptation in southern Africa: Addressing the needs of vulnerable communities. Stockholm Environment Institute, Stockholm, Sweden, pp. 56. Ziervogel, G. and F. Zermoglio, 2009: Climate change scenarios and the development of adaptation strategies in Africa: Challenges and opportunities. Climate Research, 40(2-3), 133-146. Zinyengere, N., O. Crespo, and S. Hachigonta, 2013: Crop response to climate change in Southern Africa: A comprehensive review. Global and Planetary Change, Accepted for publication 15 August 2013. Zinyowera, M.C., B.P. Jallow, R.S. Maya, H.W.O. Okoth-Ogando, L.F. Awosika, E.S. Diop, T.E. Downing, M. El- Raey, D. Le Sueur, C.H.D. Magadza, S. Toure, and C. Vogel, 1997: Africa. In: The Regional Impacts of Climate Change: An Assessment of Vulnerability. A Special Report of IPCC Working Group II. [Watson, R.T., M.C. Zinyowera, R.H. Moss, and D.J. Dokken (eds.)]. Cambridge University Press, Cambridge, UK, pp. 29-84. Zitholele Consulting, 2009: Report on the city of Umhlathuze climate change vulnerability assessment. Zitholele Consulting, Durban, South Africa, pp. 57. Subject to Final Copyedit 99 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 22-1: Major conclusions from previous IPCC assessments. Report Major conclusions Reference IPCC special Sensitivity of water resources and coastal zones to climatic parameters Zinyowera report on Identification of climate change as an additional burden on an already stressful situation et al., regional climate Major challenges for Africa: Lack of data on energy sources; uncertainties linked to 1997 change climate change scenarios (mainly for precipitation); need for integrated studies; and the necessary links between science and decision makers Third Impacts of climate change on and vulnerability of six sectors: water resources; food Desanker Assessment security; natural resources and biodiversity management; health; human settlements and et al., Report (TAR) infrastructure; desertification 2001 Adaptation strategies for each of the sectors Threats of desertification and droughts to the economy of the continent Suggestion of adaptation options: mainly linked with better resource management Identification of research gaps and needs: capacity building; data needs; development of integrated analysis; consideration of literature in other languages Fourth Vulnerability of Africa mainly due to its low adaptive capacity Boko et Assessment Sources of vulnerability mainly socioeconomic causes (demographic growth, al., 2007 Report (AR4) governance, conflicts, etc.) Impacts of climate change on various sectors: energy, tourism and coastal zones considered separately Potential impacts of extreme weather events (droughts and floods) Adaptation costs Need for mainstreaming climate change adaptation into national development policies Two case studies: 1. Food security: climate change could affect the three main components of food security 2. Traditional Knowledge: African communities have prior experience with climate variability, although this knowledge will not be sufficient to face climate change impacts. Research needs: better knowledge of climate variability; more studies on the impacts of climate change on water resources, energy, biodiversity, tourism, and health; the links between different sectors (e.g., between agriculture, land availability, and biofuels); developing links with the disaster reduction community; increasing interdisciplinary analysis of climate change; and strengthening institutional capacities Table 22-2: Under-nourishment in Africa, by number and % of total population. Undernourished 1990 1992 1999 2001 2004 2006 2007 2009 2010 2012 Million 175 205 210 220 239 (%) of total population 27.3% 25.3% 23.1% 22.6% 22.9% Source: IFAD et al., 2012 Subject to Final Copyedit 100 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 22-3: Examples of detected changes in species, natural ecosystems, and managed ecosystems in Africa that are both consistent with a climate change signal and published since the AR4. Confidence in detection of change is based on the length of study, and the type, amount and quality of data in relation to the natural variability in the particular species or system. Confidence in the role of climate being a major driver of the change is based on the extent to which the detected change is consistent with that expected under climate change, and to which other confounding or interacting non-climate factors have been considered and been found insufficient to explain the observed change. Confidence Type of Confidence in in the role change and Time scale of Potential climate Examples the detection of climate nature of observations change driver(s) of change vs. other evidence drivers Changes in Across sub-Saharan Africa, 57% ~ 25 years medium Increasing CO2, low ecosystem increase in agricultural areas and (1975-2000) changing types 15% increase in barren (largely precipitation Robust desert) areas was accompanied by patterns, increasing evidence 16% decrease in total forest cover temperatures and 5% decrease in total non-forest cover (Brink and Eva, 2009) On Mt. Kilimanjaro, increased ~ 25 years high Increasing Low vulnerability to anthropogenic fires (1976-2000) temperatures, has driven 9% decreases in montane decreasing forest and 83% decreases in precipitation subalpine forest (Hemp, 2009) In the Democratic Republic of the ~ 10 years high None proposed low Congo, total forest cover declined (2000-2010) by 2.3%, with most losses in secondary humid forest (Potapov et al., 2012) Dieback of seaward edge of ~ 35 years high Sea level rise medium mangroves in Cameroon at rates up (1975-2010) to 3 m year-1 (Ellison and Zhou, 2012) Across western Africa, central ~ 20 years high None proposed low Africa and Madagascar, net (1990-2010) deforestation was 0.28% year-1 for 1990-2000 and 0.14% year-1 for 2000-2010 (Mayaux et al., 2013) Subject to Final Copyedit 101 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Changes in Surveys of coral reefs in northern ~ 9 years high None proposed low ecosystem Tanzania indicate relative stability in (1996-2005) structure the abundance and diversity of Robust species, despite climate and non- evidence climate stressors (McClanahan et al., 2009) Analysis of sediment cores from ~ 100 years high Increasing Low Lake Victoria indicates current (1900-2000) temperatures community structure (i.e., dominated by cyanobacteria and invasive fish) was established rapidly, during the 1980s (Hecky et al., 2010) Long-term declines in density of ~ 20-50 years high Drought stress low trees and shrubs in the Sahel zone of (Senegal, 1976- induced by Senegal (Vincke et al., 2010) and 1995; Mali, 1952- decreasing Mali (Ruelland et al., 2011) 2003) precipitation Southward shift in the Sahel, Sudan, ~ 40-50 years medium Increasing medium and Guinean savanna vegetation (density,1954- temperatures, zones inferred from declines in tree 2002; diversity, decreasing density in Senegal and declines in 1960-2000) precipitation tree species richness and changes in species composition in Mauritania, Mali, Burkina Faso, Niger, and Chad (Gonzales et al., 2012) Long-term increase in shrub and tree ~67 years high Increasing CO2 low cover across mesic savanna sites (1937-2004) (700-1000 mm MAP) with contrasting land-use histories in South Africa (Wigley et al., 2009; Wigley et al., 2010) In long-term field experiments in ~ 30 50 years high In mesic site, Medium South Africa where disturbance from (1980-2010 for increasing CO2; but fire and herbivory was controlled, 600 mm MAP site; lack of response in density of trees and shrubs increased 1954-2004 for 550 semiarid site in mesic savannas (600 and 750 mm & 750 mm MAP surprising and MAP) but showed no change in a sites) unexplained semi-arid savanna (550 mm MAP) (Buitenwerf et al., 2012) Changes in A reconstruction of drought history ~ 550 years (1456 high Increasing low ecosystem in Tunisia and Algeria based on tree 2002) temperatures, physiology ring records from Cedrus atlantica decreasing Moderate and Pinus halepensis indicates that a precipitation evidence 1999-2002 drought was the most severe since the 15th century (Touchan et al., 2008) Across 79 African tropical forest ~ 40 years high Increasing CO2 Medium plots, above-ground carbon storage (1968-2007) in live trees increased by 0.63 Mg C ha-1 yr-1 (Lewis et al., 2009) Increased stratification and reduced ~ 90 years high Increasing High nutrient fluxes and primary (1913-2000) temperatures productivity in Lake Tanganyika (Verburg and Hecky, 2009) Recent increases in surface ~ 1,500 years high Increasing High temperatures and decreases in (500-2000) temperatures productivity of Lake Tanganyika exceed the range of natural variability (Tierney et al., 2010) Subject to Final Copyedit 102 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Changes in species The range of Aloe dichotoma, a Namib ~ 100 high Increasing Medium distributions, Desert tree, is shifting poleward, but years temperatures, physiology, or extinction along trailing edge exceeds (1904- decreasing behavior colonization along leading edge (Foden et 2002) precipitation Moderate evidence al., 2007) On Tsaratanana Massif, the highest ~ 10 high Increasing Medium mountain in Madagascar, reptiles and years temperatures amphibians are moving upslope (1993- (Raxworthy et al., 2008) 2003) Pomacentrus damselfish species vary in minutes high Increasing CO2 low avoidance of predation-related mortality to days under elevated CO2 (Ferrari et al., 2011) (Nov.- Dec. 2009) In greenhouse experiments, growth of ~1-2 high Increasing CO2 medium seedlings of woody savanna species years (Acacia karoo and Terminalia sericea) was enhanced at elevated CO2 (Bond and Midgley, 2012) Table 22-4: Projected changes in agro-climatic suitability for perennial crops in Africa by mid-century under an A2 scenario. Crop Suitability change Country Source Coffee Increased suitability at high latitudes; decreased Kenya Läderach et al., 2010 suitability at low latitudes Tea Decreased suitability Uganda Eitzinger et al., Increased suitability at high latitudes; decreased Kenya 2011a,b suitability at low latitudes Cocoa Constant or increased suitability at high Ghana, Côte d Ivoire Läderach et al., 2011c latitudes; decreased suitability at low latitudes Cashew Increased suitability Ghana, Côte d Ivoire Läderach et al., 2011a Cotton Decreased suitability Ghana, Côte d Ivoire Läderach et al., 2011b Subject to Final Copyedit 103 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 22-5: Cross-cutting approaches for equity and social justice in adaptation. Equitable Key issues to address for Factors that could adaptation Opportunities Lessons learned adaptation cause maladaptation approach Gender Lack of empowerment and Employment Women s aptitude for Security of tenure mainstreamed participation in decision- opportunities not long-term thinking, over land and adaptation in Africa making (Patt et al., 2009) sufficiently extended trusting and integrating resource access is Climate impacts increase to women in scientific knowledge, and critical for enabling women s household roles, adaptation initiatives taking decisions under enhanced adaptive with risk of girls missing (Madzwamuse, 2010) uncertainty (Patt et al., capacity of women school to assist (Raworth, Failure to incorporate 2009) (ADF, 2010) 2008; Romero González et power relations in Potential long-term Research on al., 2011; UNDP, 2011b) adaptation responses increase in women s understanding Male adaptation strategies (Djoudi and empowerment and social different adaptive e.g. migration risk Brockhaus, 2011; and economic status strategies of benefit increasing women s Romero González et (Djoudi and Brockhaus, for women and men vulnerability (Djoudi and al., 2011) 2011) is needed Brockhaus, 2011) Women opportunistically using development projects for adaptation (Nielsen, 2010) Child-centered Children and youth Limits to children s Using approaches that Positive role of approaches to represent over 60% of agency related to stress agency and children and youth adaptation Africa s population, yet power imbalances empowerment, and as change agents for their issues are largely between children and innovative energies of climate adaptation, absent from adaptation adults, and different youth; build on targeted within appropriate policy (ADF, 2010) cultural contexts adaptation initiatives, such enabling Children s differential (Seballos et al., 2011) as child-centred disaster environment vulnerability to projected risk reduction and Child-sensitive climate impacts is high, adaptation (ADF, 2010; programmes and particularly to hunger, Seballos et al., 2011) policies can reduce malnutrition and disasters risks children face (UNICEF, 2007) from disasters (Seballos et al, 2011) Funding for climate resilience programmes will protect children s basic rights (UNICEF, 2010; UNICEF, 2011) Human rights-based Common critical rights Lack of recognition Using the HRBA lens to Applying HRBA approaches (HRBA) issues for local and promotion of their understand climate risk presents a communities are human rights blocks necessitates risk analysis framework for land/resource rights, indigenous peoples to probe the root causes of addressing gender equality, and coping and adaptation differential disaster risk conflicting rights political voice and fair capacities (UNPFII, vulnerabilities, to enable and interests, adjudication of grievances 2008) structural, sustainable necessary for for the poor and excluded responses (Urquhart, building resilience (Castro et al., 2012) 2014) and equitable adaptation responses (Nilsson and Schnell, 2010) Subject to Final Copyedit 104 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 22-6: Key risks from climate change and the potential for risk reduction through mitigation and adaptation in Africa. Key risks are identified based on assessment of the literature and expert judgments made by authors of the various WGII AR5 chapters, with supporting evaluation of evidence and agreement in the referenced chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels are presented for the near-term era of committed climate change (here, for 2030-2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the longer-term era of climate options (here, for 2080-2100), for global mean temperature increase of 2°C and 4°C above preindustrial levels. For each timeframe, risk levels are estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols. Subject to Final Copyedit 105 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 22-7: Land inundated and economic impacts in Cape Town based on a risk assessment (Cartwright, 2008). Sea level rise scenarios Land inundated Economic impacts (for 25 years) Scenario 1 (+ 2.5 to 6.5 m depending 25.1 km2 (1% of the total CT area) 5.2 billion R (794 million US$) on the exposure) 95% Scenario 2 (+ 4.5 m) 85% 60.9 km2 (2% of the total CT area) 23.7 billion R (30.3 billion US$) Scenario 3 (+6.5 m) 20% 95 km2 (4% of the total CT area) 54.8 billion R Note: The economic impacts are determined based on the value of properties, losses of touristic revenues and the cost of infrastructure replacement. The total geographical gross product for Cape Town in 2008 was 165 billion of Rands. Table 22-8: Research gaps in different sectors. Key Sectors Gaps observed Climate Science Research in climate and climate impacts would be greatly enhanced if data custodians and researchers worked together to use observed station data in scientific studies. Research into regional climate change and climate impacts relies on observed climate and hydrological data as an evaluative base. These data are most often recorded by meteorological institutions in each country and sold to support data collection efforts. However, African researchers are generally excluded from access to these critical data due to the high costs involved which hinders both climate and climate impacts research. Downscaling GCM data to the regional scale captures the influence of topography on the regional climate. Regional climate information is essential for understanding regional climate processes, regional impacts and potential future changes in these. Additionally, impacts models such as hydrology and crop models generally require input data at a resolution higher than what GCMs can provide. Regional downscaling, either statistically or through using regional climate models, can provide information at these scales and can also change the sign of GCM-projected rainfall change over topographically complex areas [22.2.2.2]. Ecosystems Monitoring networks for assessing long-term changes to critical ecosystems such as coastal ecosystems, lakes, mountains, grasslands, forests, wetlands, deserts, and savannas to enhance understanding of long- term ecological dynamics, feedbacks between climate and ecosystems, the effects of natural climate variability on ecosystems, the limits of natural climate variability, and the marginal additional effects of global climate forcing. Develop the status of protected areas to include climate change effects Food Systems Socioeconomic and environmental tradeoffs of biofuel production, especially the effect on land use change and food and livelihood security; better agronomic characterization of biofuel crops to avoid maladaptive decisions with respect to biofuel production Vulnerability to and impacts of climate change on food systems (production, transport, processing, storage, marketing and consumption) Impacts of climate change on urban food security, and dynamic of rural-urban linkages in vulnerability and adaptive capacity Impacts of climate change on food safety and quality Water resources Characterization of Africa s groundwater resource potential; understanding interactions between non- climate and climate drivers as related to future groundwater resources. Impacts of climate change on water quality, and how this links to food and health security Decision making under uncertainty with respect to water resources given limitations of climate models for adequately capturing future rainfall projections Subject to Final Copyedit 106 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Human security Research to explore and monitor the links between climate change and migration and its potential and urban areas negative effects on environmental degradation; the potential positive role of migration in climate change adaptation. Improved methods and research to analyse the relation between climate change and violent conflict. Livelihoods and Methodologies for cyclical learning and decision-support to enable anticipatory adaptation in contexts of poverty high poverty and vulnerability (Tschakert and Dietrich, 2010) Frameworks to integrate differentiated views of poverty into adaptation and disaster risk reduction, and to better link these with social protection in different contexts Ethical and political dimensions of engaging with local and traditional knowledge on climate change Health Research and improved methodologies (including longitudinal studies) to assess and quantify the impact of climate change on vector borne, foodborne, waterborne, nutrition, heatstress and indirect impacts on HIV. Research to quantify the direct and indirect health impacts of extreme weather events in Africa; injuries, mental illness; health infrastructure Frameworks and research platforms to be developed with other sectors to determine how underlying risks (for example food security) will be addressed to improve health outcomes. Adaptation Research to develop home-grown and to localize global adaptation technologies to build resilience Equitable adaptation frameworks to deal with high uncertainty levels and integrate marginalized groups; and that identify and eliminate multi-level constraints to women s adaptive ability Multi--tiered approach to building institutional and community capacity to respond to climate risk Potential changes in economic and social systems under different climate scenarios, to understand the implications of adaptation and planning choices (Clements et al., 2011) Principles/determining factors for effective adaptation, including community-based adaptation Understanding synergies and trade-offs between different adaptation and mitigation approaches (Chambwera and Anderson, 2011) Additional national and sub-natinal modeling and analysis of the economic costs of impacts and adaptation, including of the soft costs of impacts and adaptation Monitoring adaptation Other Methods in vulnerability analysis for capturing the complex interactions in systems across scales Understanding compound impacts from concomitant temperature and precipitation stress, e.g. effect on a particular threshold of a heatwave occurring during a period of below normal precipitation. Subject to Final Copyedit 107 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 22-1: Observed and simulated variations in past and projected future annual average precipitation and temperature. Observed differences in the Climate Research Unit, University of East Anglia data (CRU) are shown between the 1986-2005 and 1906-1925 periods, with white indicating areas where the difference between the 1986- 2005 and 1906-1925 periods is less than twice the standard deviation of the 20 20-year periods beginning in the years 1906 through 1925. For CMIP5, white indicates areas where <66% of models exhibit a change greater than twice the baseline standard deviation of the respective model s 20 20-year periods ending in years 1986 through 2005. Grey indicates areas where >66% of models exhibit a change greater than twice the respective model baseline standard deviation, but <66% of models agree on the sign of change. Colors with circles indicate the ensemble-mean change in areas where >66% of models exhibit a change greater than twice the respective model baseline standard deviation and >66% of models agree on the sign of change. Colors without circles indicate areas where >90% of models exhibit a change greater than twice the respective model baseline standard deviation and >90% of models agree on the sign of change. The realizations from each model are first averaged to create baseline-period and future-period mean and standard deviation for each model, from which the multi-model mean and the individual model signal-to-noise ratios are calculated. The baseline period is 1986-2005. The late-21st Century period is 2081- 2100. The mid-21st century period is 2046-2065. Subject to Final Copyedit 108 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 109 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 22-2: Observed and simulated variations in past and projected future annual average temperature over EAC- IGAD-Egypt, ECCAS, ECOWAS, SADC and UMA. Black lines show various estimates from observational measurements. Shading denotes the 5-95 percentile range of climate model simulations driven with "historical" changes in anthropogenic and natural drivers (63 simulations), historical changes in "natural" drivers only (34), the "RCP2.6" emissions scenario (63), and the "RCP8.5" (63). Data are anomalies from the 1986-2005 average of the individual observational data (for the observational time series) or of the corresponding historical all-forcing simulations. Further details are given in Box 21-3. [Illustration to be redrawn to conform to IPCC publication specifications.] Figure 22-3: Left: Confidence in detection and in attribution of observed climate change over Africa to anthropogenic emissions. All detection assessments are against a reference of no change, while all attribution assessments concern a major role of anthropogenic emissions in the observed changes. See 22.2, and SREX-3, and WGI AR5 10 for details. Right: Confidence in detection and in attribution of the impacts of observed regional climate change on various African systems. All detection assessments are against a reference of no change, except "Kenyan Highlands malaria" (changes due to vaccination, drug resistance, demography, and livelihoods), "Great Lakes fisheries" (changes due to fisheries management and land use) and Adapting South African farmers (economic changes). Attribution is to a major role or a minor role of observed climate change, as indicated. See 22.2.2, 22.3.2.1, 22.3.2.2, 2.3.3, 22.4.2, 22.3.4.4, 22.3.5.4, 22.4.5.7 and Tables 18-5, 18-6, 18-7, and 18-9 for details. Assessments follow the methods outlined in 18.2. Subject to Final Copyedit 110 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 22-4: Left Projected biome change from the periods 1961-1990 to 2071-2100 using the MC1 Dynamic Vegetation Model. Change is indicated if any of nine combinations of three GCMs (CSIRO Mk3, HadCM3, MIROC 3.2 medres) and three emissions scenarios (B1, A1B, A2) project change and is thus a worst-case scenario. Colours represent the future biome predicted. Right Vulnerability of ecosystems to biome shifts based on historical climate (1901-2002) and projected vegetation (2071-2100), where all nine GCM-emissions scenario combinations agree on the projected biome change. Source: Gonzalez et al. (2010). Subject to Final Copyedit 111 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 22-5: The effect of rainfall and temperature changes on mean crop yield. Mean crop yield change (%) relative to the 1961 90 baseline for 7 temperatures (x-axis) and 5 rainfall (y-axis) scenarios. Results are shown as the average over the 35 stations across West Africa and the 6 cultivars of sorghum and millet. White triangles and circles are the projected anomalies computed by several CMIP3 GCMs and three IPCC emission scenarios (B1, A1B, A2) for 2071 90 and 2031 50, respectively. Projections from CMIP5 GCMs and three RCPs (4.5, 6.0 and 8.5) are represented by grey triangles and circles. Models and scenarios names are displayed in figure S2 (available at stacks.iop.org/ERL/8/014040/mmedia). Past observed climate anomalies from CRU data are also projected by computing 10-year averages (e.g. '1940' is for 1941 50). All mean yield changes are significant at a 5% level except boxes with a diagonal line. Source: Sultan et al., 2013. Subject to Final Copyedit 112 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 22-6: Total additional costs of adaptation per year from 2000 to 2100 for Tanzania (including beach nourishment and sea and river dikes). The values do not consider the existing adaptation deficit (values in $US2005), without discounting. Source: Kebede et al., 2010. Subject to Final Copyedit 113 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure UP-1: Upper panel: Schematic hypothetic mechanism of increasing coastal wind-driven upwelling at eastern boundary systems, where differential warming rates between land and ocean results in increased land-ocean pressure gradients (1) that produce stronger alongshore winds (2) and offshore movement of surface water through Ekman transport (3), and increased upwelling of deep cold nutrient rich waters to replace it (4). Lower panel: potential consequences of climate change in upwelling systems. Increasing stratification and uncertainty in wind stress trends result in uncertain trends in upwelling. Increasing upwelling may result in higher input of nutrients to the euphotic zone, and increased primary production, which in turn may enhance pelagic fisheries, but also decreased coastal fisheries due to an augmented exposure of coastal fauna to hypoxic, low pH waters. Decreased upwelling may result in lower primary production in these systems with direct impacts on pelagic fisheries productivity. Subject to Final Copyedit 114 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 22 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure WE-1: The water-energy-food nexus as related to climate change. The interlinkages of supply/demand, quality and quantity of water, energy and food/feed/fiber with changing climatic conditions have implications for both adaptation and mitigation strategies. Subject to Final Copyedit 115 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 23. Europe Coordinating Lead Authors Sari Kovats (UK), Riccardo Valentini (Italy) Lead Authors Laurens M. Bouwer (Netherlands), Elena Georgopoulou (Greece), Daniela Jacob (Germany), Eric Martin (France), Mark Rounsevell (UK), Jean-Francois Soussana (France) Contributing Authors Martin Beniston (Switzerland), Maria Vincenza Chiriaco (Italy), Philippe Cury (France), Michael Davies (UK), Paula Harrison (UK), Olaf Jonkeren (Italy), Mark Koetse (Netherlands), Markus Lindner (Finland), Andreas Matzarakis (Germany), Reinhard Mechler (Germany), Annette Menzel (Germany), Marc Metzger (UK), Luca Montanarella (Italy), Antonio Navarra (Italy), Juliane Peterson (Germany), Martin Price (UK), Boris Revich (Russian Federation), Piet Rietveld (Netherlands), Cristina Sabbioni (Italy), Yannis Sarafidis (Greece), Philipp Schmidt-Thomé (Finland), Vegard Skirbekk (Austria), Donatella Spano (Italy), Jan E. Vermaat (Netherlands), Paul Watkiss (UK), Meriwether Wilson (UK), Thomasz Zylicz (Poland) Review Editors Lucka Kajfez Bogataj (Slovenia), Roman Corobov (Moldova), Ramón Vallejo (Spain) Contents Executive Summary 23.1. Introduction 23.1.1. Scope and Route Map of Chapter 23.1.2. Policy Frameworks 23.1.3. Conclusions from Previous Assessments 23.2. Current and Future Trends 23.2.1 Non- Climate Trends 23.2.2. Observed and Projected Climate Change 23.2.2.1. Observed Climate Change 23.2.2.2. Projected Climate Changes 23.2.2.3. Projected Changes in Climate Extremes 23.2.3. Observed and Projected Trends in the Riverflow and Drought 23.3. Implications of Climate Change for Production Systems and Physical Infrastructure 23.3.1. Settlements 23.3.1.1. Coastal Flooding 23.3.1.2. River and Pluvial Flooding 23.3.1.3. Windstorms 23.3.1.4.Mass Movements and Avalanches 23.3.2. Built Environment 23.3.3. Transport 23.3.4. Energy Production, Transmission, and Use 23.3.5. Industry and Manufacturing 23.3.6. Tourism 23.3.7. Insurance and Banking 23.4. Implications of Climate Change for Agriculture, Fisheries, Forestry, and Bioenergy Production 23.4.1. Plant (Food) Production Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 23.4.2. Livestock Production 23.4.3. Water Resources and Agriculture 23.4.4. Forestry 23.4.5. Bioenergy Production 23.4.6. Fisheries and Aquaculture 23.5. Implications of Climate Change for Health and Social Welfare 23.5.1. Human Population Health 23.5.2. Critical Infrastructure 23.5.3. Social Impacts 23.5.4. Cultural Heritage and Landscapes 23.6. Implications of Climate Change for the Protection of Environmental Quality and Biological Conservation 23.6.1. Air Quality 23.6.2. Soil Quality and Land Degradation 23.6.3. Water Quality 23.6.4. Terrestrial and Freshwater Ecosystems 23.6.5. Coastal and Marine Ecosystems 23.7. Cross-Sectoral Adaptation Decision-making and Risk Management 23.7.1. Coastal Zone Management 23.7.2. Integrated Water Resource Management 23.7.3. Disaster Risk Reduction and Risk Management 23.7.4. Land Use Planning 23.7.5. Rural Development 23.7.6. Economic Assessments of Adaptation 23.7.7. Barriers and Limits to Adaptation 23.8. Co-Benefits and Unintended Consequences of Adaptation and Mitigation 23.8.1. Production and Infrastructure 23.8.2. Agriculture, Forestry, and Bioenergy 23.8.3. Social and Health Impacts 23.8.4. Environmental Quality and Biological Conservation 23.9. Synthesis of Key Findings 23.9.1. Key Vulnerabilities 23.9.2. Climate Change Impacts Outside Europe and Inter-Regional Implications 23.9.3. Effects of Observed Climate Change in Europe 23.9.4. Key Knowledge Gaps and Research Needs References Chapter Boxes 23-1. Assessment of Climate Change Impacts on Ecosystem Services by Sub-Region 23-2. Implications of Climate Change for European Wine and Vineyards 23-3. National and Local Adaptation Strategies Frequently Asked Questions 23.1: Will I still be able to live on the coast in Europe? 23.2: Will climate change introduce new infectious diseases into Europe? 23.3: Will Europe need to import more food because of climate change? Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Executive Summary Observed climate trends and future climate projections show regionally varying changes in temperature and rainfall in Europe [high confidence] [23.2.2], in agreement with AR4 findings, with projected increases in temperature throughout Europe and increasing precipitation in Northern Europe and decreasing precipitation in Southern Europe [23.2.2.2]. Climate projections show a marked increase in high temperature extremes [high confidence], meteorological droughts [medium confidence] [23.2.3] and heavy precipitation events [high confidence] [23.2.2.3] with variations across Europe, and small or no changes in wind speed extremes [low confidence] except increases in winter wind speed extremes over Central and Northern Europe [medium confidence] [23.2.2.3]. Observed climate change in Europe has had wide ranging effects throughout the European region including: the distribution, phenology, and abundance of animal, fish and plant species [high confidence] [23.6.4, Table 23.6]; stagnating wheat yields in some sub-regions [medium confidence, limited evidence] [23.4.1]; and forest decline in some sub-regions [medium confidence] [23.4.4]. Climate change has affected both human health (from increased heat waves) [medium confidence] [23.5.1] and animal health (changes in infectious diseases) [high confidence] 23.4.5]. There is less evidence of impacts on social systems attributable to observed climate change, except in pastoralist populations [low confidence]. Climate change will increase the likelihood of systemic failures across European countries caused by extreme climate events affecting multiple sectors [medium confidence] [23.2.2.3, 23.2.3, 23.3, 23.4, 23.5, 23.6, 23.9.1]. Extreme weather events currently have significant impacts in Europe in multiple economic sectors as well as adverse social and health effects [high confidence] [Table 23.1]. There is limited evidence that resilience to heat waves and fires has improved in Europe [medium confidence] [23.9.2, 23.5.], while some countries have improved their flood protection following major flood events [23.9.2, 23.7.3]. Climate change is very likely to increase the frequency and intensity of heat waves, particularly in Southern Europe [high confidence] [23.2.2] with mostly adverse implications for health, agriculture, forestry, energy production and use, transport, tourism, labour productivity, and the built environment [Table 23-1, 23.3.2, 23.3.3, 23.3.4, 23.3.6, 23.4.1, 23.4.2, 23.4.3, 23.4.4, 23.5.1]. The provision of ecosystem services is projected to decline across all service categories in response to climate change in Southern Europe and Alpine sub-regions [high confidence] [23.9.1, Box 23-1]. Both gains and losses in the provision of ecosystem services are projected for the other European sub-regions [high confidence], but the provision of cultural services is projected to decline in the Continental, Northern and Southern sub-regions [low confidence] [Box 23-1]. Climate change is expected to impede economic activity in Southern Europe more than in other sub-regions [medium confidence] [Table 23.4, 23.9.3], and may increase future intra-regional disparity [low confidence] [23.9.3]. There are also important differences in vulnerability within sub-regions, for example, plant species and some economic sectors are most vulnerable in high mountain areas due to lack of adaptation options [medium confidence][23.9.1.]. Southern Europe is particularly vulnerable to climate change [high confidence] as multiple sectors will be adversely affected (tourism, agriculture, forestry, infrastructure, energy, population health) [high confidence] [23.9] [Box 23-3]. The impacts of sea level rise on populations and infrastructure in coastal regions can be reduced by adaptation [medium confidence] [23.3.1, 23.5.3]. Populations in urban areas are particularly vulnerable to climate change impacts due to the high density of people and built infrastructure [medium confidence] [23.3, 23.5.1]. Synthesis of evidence across sectors and sub-regions confirm that there are limits to adaptation from physical, social, economic and technological factors [high confidence] [23.5]. Adaptation is further impeded because climate change affects multiple sectors [23.10]. The majority of published assessments are based on climate projections in the range 1-4 degrees global mean temperature per century. Limited evidence exists regarding the potential impacts in Europe under high rates of warming (>4 degrees global mean temperature per century) [23.9.1]. Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Impacts by Sector Sea level rise and increases in extreme rainfall are projected to further increase coastal and river flood risk in Europe and, without adaptive measures, will substantially increase flood damages (people affected and economic losses) [high confidence] [23.3.1, 23.5.1]. Adaptation can prevent most of the projected damages [high confidence based on medium evidence, high agreement] but there may be constraints to building flood defences in some areas [23.3.1, 23.7.1, 23.8.3]. Direct economic river flood damages in Europe have increased over recent decades [high confidence] but this increase is due to development in flood zones and not due to observed climate change [23.3.1.2, SREX 4.5]. Some areas in Europe show changes in river flood occurrence related to observed changes in extreme river discharge [medium confidence] [23.2.3]. Climate change is projected to affect the impacts of hot and cold weather extremes on transport leading to economic damage and/or adaptation costs, as well as some benefits (e.g. reduction of maintenance costs) during winter [medium confidence] [23.3.3]. Climate change is projected to reduce severe accidents in road transport [medium confidence] and adversely affect inland water transport in summer in some rivers (e.g. the Rhine) after 2050 [medium confidence]. Damages to rail infrastructure from high temperatures may also increase [medium confidence]. Adaptation through maintenance and operational measures can reduce adverse impacts to some extent. Climate change is expected to affect future energy production and transmission [23.3.4]. Hydropower production is likely to decrease in all sub-regions except Scandinavia [high confidence] [23.3.4]. Climate change is unlikely to affect wind energy production before 2050 [medium confidence] but will have a negative impact in summer and a varied impact in winter after 2050 [medium confidence]. Climate change is likely to decrease thermal power production during summer [high confidence] [23.3.4]. Climate change will increase the problems associated with overheating in buildings [medium confidence] [23.3.2]. Although climate change is very likely to decrease space heating demand [high confidence], cooling demand will increase [very high confidence] although income growth mostly drives projected cooling demand up to 2050 [medium confidence] [23.3.4]. More energy efficient buildings and cooling systems as well as demand-side management will reduce future energy demands [23.3.4]. After 2050, tourism activity is projected to decrease in southern Europe [low confidence] and increase in Northern and Continental Europe [medium confidence]. No significant impacts on the tourism sector are projected before 2050 in winter or summer tourism except for ski tourism in low altitude sites and under limited adaptation [medium confidence] [23.3.6]. Artificial snowmaking may prolong the activity of some ski resorts [medium confidence] [23.3.6]. Climate change is likely to increase cereal yields in Northern Europe [medium confidence, disagreement] but decrease yields in Southern Europe [high confidence] [23.4.1]. In Northern Europe, climate change is very likely to extend the seasonal activity of pests and plant diseases [high confidence] [23.4.1]. Yields of some arable crop species like wheat have been negatively affected by observed warming in some European countries since 1980s [medium confidence, limited evidence] [23.4.1] Compared to AR4, new evidence regarding future yields in Northern Europe, is less consistent regarding the magnitude and sign of change. Climate change may adversely affect dairy production in Southern Europe because of heat stress in lactating cows [medium confidence] [23.4.2]. Climate change has contributed to vector-borne disease in ruminants in Europe [high confidence] [23.4.2] and northward expansion of tick disease vectors [medium confidence] [23.4.2, 23.5.1]. Climate change will increase irrigation needs [high confidence] but future irrigation will be constrained by reduced runoff, demand from other sectors, and by economic costs [23.4.1, 23.4.3]. By 2050s, irrigation will not be sufficient to prevent damage from heat waves to crops in some sub-regions [medium confidence]. System costs will increase under all climate scenarios [high confidence] [23.4.3]. Integrated management of water, also across countries boundaries, is needed to address future competing demands between agriculture, energy, conservation and human settlements [23.7.2]. As a result of increased evaporative demand, climate change is likely to significantly reduce water availability from river abstraction and from groundwater resources [medium confidence], in the context of increased demand (from agriculture, energy and industry, and domestic use) and cross-sectoral implications which are not Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 fully understood [23.4.3, 23.9.1]. Some adaptation is possible through uptake of more water efficient technologies and water saving strategies [23.4.3, 23.7.2, 23.9.1]. Climate change will change the geographic distribution of wine grape varieties [high confidence] and this will reduce the value of wine products and the livelihoods of local wine communities in Southern and Continental Europe [medium confidence] and increase production in Northern Europe [low confidence] [23.4.1, 23.3.5, 23.5.4, Box 23-2]. Some adaptation is possible through technologies and good practice [Box 23-2]. Climate warming will increase forest productivity in northern Europe [medium confidence] [23.4.4], although damage from pests and diseases in all sub-regions will increase due to climate change [high confidence] [23.4.4]. Wildfire risk in Southern Europe [high confidence] and damages from storms in central Europe [low confidence] may also increase due to climate change [23.4.4]. Climate change is likely to cause ecological and socio-economic damages from shifts in forest tree species range (from south-west to north-east) [medium confidence], and in pest species distributions [low confidence] [23.4.4]. Forest management measures can enhance ecosystem resilience [medium confidence] [23.4.4]. Observed warming has shifted marine fish species ranges to higher latitudes [high confidence] and reduced body size in species [medium confidence] [23.4.6]. There is limited and diverging evidence on climate change impacts on net fisheries economic turnover. Local economic impacts attributable to climate change will depend on the market value of (high temperature tolerant) invasive species [23.4.6]. Climate change is unlikely to entail relocation of fishing fleets [high confidence] [23.4.6]. Observed higher water temperatures have adversely affected both wild and farmed freshwater salmon production in the southern part of their distribution [high confidence] [23.4.6]. High temperatures may increase the frequency of harmful algal blooms [low confidence] [23.4.6]. Climate change will affect bioenergy cultivation patterns in Europe by shifting northward their potential area of production [medium confidence] [23.4.5]. Elevated atmospheric CO2 can improve drought tolerance of bioenergy crop species due to improved plant water use, maintaining high yields in future climate scenarios in temperate regions [low confidence] [23.4.5]. Climate change is likely to affect human health in Europe. Heat-related deaths and injuries are likely to increase, particularly in Southern Europe [medium confidence] [23.5.1]. Climate change may change the distribution and seasonal pattern of some human infections, including those transmitted by arthropods [medium confidence], and increase the risk of introduction of new infectious diseases [low confidence] [23.5.1]. Climate change and sea level rise may damage European cultural heritage, including buildings, local industries, landscapes, archaeological sites, and iconic places [medium confidence] and some cultural landscapes may be lost forever [low confidence] [23.5.4] [Table 23.3]. Climate change may adversely affect background levels of tropospheric ozone [low confidence, limited evidence, low agreement], assuming no change in emissions, but the implications for future particulate pollution (which is more health-damaging) are very uncertain [23.6.1]. Higher temperatures may have affected trends in ground level tropospheric ozone [low confidence] [23.6.1.]. Climate change is likely to decrease surface water quality due to higher temperatures and changes in precipitation patterns [medium confidence] [23.6.3], and is likely to increase soil salinity in coastal regions [low confidence] [23.6.2]. Climate change may also increase soil erosion (from increased extreme events) and reduce soil fertility [low confidence, limited evidence] [23.6.2]. Observed climate change is affecting a wide range of flora and fauna, including plant pests and diseases [high confidence] [23.4.1, 23.4.4] and the disease vectors and hosts [medium confidence] [23.4.3]. Climate change is very likely to cause changes in habitats and species, with local extinctions [high confidence] and continental scale shifts in species distributions [medium confidence] [23.6.4]. The habitat of alpine plants is very likely to be significantly reduced [high confidence] [23.6.4]. Phenological mismatch will constrain both terrestrial and marine ecosystem functioning under climate change [high confidence] [23.6.4, 23.6.5], with a reduction in some ecosystem services [low confidence] [23.6.4, Box 23-1]. The introduction and expansion of invasive species, especially those with high migration rates, from outside Europe is likely to increase with climate change [medium confidence] Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [23.6.4]. Climate change is likely to entail the loss or displacement of coastal wetlands [high confidence] [23.6.5]. Climate change threatens the effectiveness of European conservation areas [low confidence] [23.6.4], and stresses the need for habitat connectivity through specific conservation policies [23.6.4]. Adaptation The capacity to adapt in Europe is high compared to other world regions, but there are important differences in impacts and in the capacity to respond between and within the European sub-regions. In Europe, adaptation policy has been developed at international (European Union), national and local government level [23.7], including the prioritisation of adaptation options. There is limited systematic information on current implementation or effectiveness of adaptation measures or policies [Box 23-3]. Some adaptation planning has been integrated into coastal and water management, as well as disaster risk management [23.7.1, 23.7.2, 23.7.3]. There is limited evidence of adaptation planning in rural development or land-use planning [23.7.4, 23.7.5]. Adaptation will incur a cost, estimated from detailed bottom-up sector-specific studies for coastal defences, energy production, energy use, and agriculture [23.7.6]. The costs of adapting buildings (houses, schools, hospitals) and upgrading flood defences increase under all scenarios relative to no climate change [high confidence] [23.3.2]. Some impacts will be unavoidable due to limits (physical, technological, social, economic or political) [Table 23-3, 23.7.7]. There is also emerging evidence regarding opportunities and unintended consequences of policies, strategies and measures that address adaptation and/or mitigation goals [23.8]. Some agricultural practices can reduce GHG emissions and also increase resilience of crops to temperature and rainfall variability [23.8.2]. There is evidence for unintended consequences of mitigation policies in the built environment (especially dwellings) and energy sector [medium confidence] [23.8.1]. Low carbon policies in the transport and energy sectors to reduce emissions are associated with large benefits to human health [23.8.3] [high confidence]. 23.1. Introduction This chapter reviews the scientific evidence published since AR4 on observed and projected impacts of anthropogenic climate change in Europe and adaptation responses. The geographical scope of this chapter is the same as in AR4 with the inclusion of Turkey. Thus, the European region includes all countries from Iceland in the west to Russia (west of the Urals) and the Caspian Sea in the east, and from the northern shores of the Mediterranean and Black Seas and the Caucasus in the south to the Arctic Ocean in the north. Impacts above the Arctic Circle are addressed in the Polar Regions Chapter 28 and impacts in the Baltic and Mediterranean Seas are addressed in the Open Oceans Chapter 30. Impacts in Malta, Cyprus, and other island states in Europe are discussed in the Small Island Chapter 29. The European region has been divided into 5 sub-regions (see Figure 23-1): Atlantic, Alpine, Southern, Northern, and Continental. The sub-regions are derived by aggregating the climate zones developed by (Metzger et al., 2005) and therefore represent geographical and ecological zones rather than political boundaries. The scientific evidence has been evaluated to compare impacts across (rather than within) sub-regions, although this is not always possible, depending on the scientific information available. [INSERT FIGURE 23-1 HERE Figure 23-1: Sub-regional classification of the IPCC Europe region. Based on Metzger et al., 2005.] Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 23.1.1. Scope and Route Map of Chapter The chapter is structured around key policy areas. Sections 23.3 to 23.6 summarise the latest scientific evidence on sensitivity climate, observed impacts and attribution, projected impacts and adaptation options, with respect to four main categories of impacts: Production systems and physical infrastructure Agriculture, fisheries, forestry and bioenergy production Health protection and social welfare Protection of environmental quality and biological conservation. The benefit of assessing evidence in a regional chapter is that impacts across sectors can be described, and interactions between impacts can be identified. Further, the cross-sectoral decision making required to address climate change can be reviewed. The chapter also includes sections that were not in AR4. As adaptation and mitigation policy develops, the evidence for potential co-benefits and unintended consequences of such strategies is reviewed (Section 23.8). The final section synthesise the key findings with respect to: observed impacts of climate change, key vulnerabilities and research and knowledge gaps. The chapter evaluates the scientific evidence in relation to the five sub-regions discussed above. The majority of the research in the Europe region is for impacts in countries in the European Union due to targeted research funding through the European Commission and national governments which means that countries in eastern Europe and Russia are less well represented in this chapter. Further, regional assessments may be reported for the EU15, EU27 or EEA (32) group of countries [Table SM23-1]. 23.1.2. Policy Frameworks Since AR4, there have been significant changes in Europe in responses to climate change. More countries now have adaptation and mitigation policies in place. An important force for climate policy development in the region is the European Union (EU). EU Member States have mitigation targets, as well as the overall EU target, with both sectoral and regional aspects to the commitments. Adaptation policies and practices have been developed at the international, national and local levels although research on implementation of such policies is limited. Due to the vast range of policies, strategies and measures it is not possible to describe them extensively here. However, adaptation in related to cross-sectoral decision-making is discussed in section 23.7 (see also Box 23-3 on national adaptation policies). The European Climate Adaptation Platform (Climate-ADAPT) catalogues adaptation actions reported by EU Member States (EC, 2013b). The EU Adaptation Strategy was adopted in 2013 (EC, 2013a). See Chapter 15 for a more extensive discussion of institutions and governance in relation to adaptation planning and implementation. 23.1.3. Conclusions from Previous Assessments AR4 documented a wide range of impacts of observed climate change in Europe (AR4 WG2 Chapter 12). The SREX confirmed increases in warm days, warm nights and decreases in cold days and cold nights since 1950 (high confidence, SREX-3.3.1). Extreme precipitation increased in part of the continent, mainly in winter over western- central Europe and European Russia (medium confidence, SREX-3.3.2). Dryness has increased mainly in Southern Europe (medium confidence, SREX-3.3.2). Climate change is expected to magnify regional differences within Europe for agriculture and forestry because water stress was projected to increase over central and southern Europe (AR4-12.4.1, SREX-3.3.2, SREX-3.5.1). Many climate-related hazard were projected to increase in frequency and intensity, but with significant variations within the region (AR4-12.4). The AR4 identified that climate changes would pose challenges to many economic sectors and was expected to alter the distribution of economic activity within Europe (high confidence). Adaptation measures were evolving from reactive disaster response to more proactive risk management. A prominent example was the implementation of heat Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 health warning systems following the 2003 heat wave event (AR4 WG2 12.6.1, SREX 9.2.1). National adaptation plans were developed and specific plans were incorporated in European and national policies (AR4 WG2 12.2.3, 12.5) but these were not yet evaluated (AR4 WG2 12.8). 23.2. Current and Future Trends 23.2.1 Non- Climate Trends European countries are diverse in both demographic and economic trends. Population health and social welfare has improved everywhere in Europe, with reductions in adult and child mortality rates, but social inequalities both within and between countries persist (Marmot et al., 2012). Population has increased in most EU27 countries, primarily due to net immigration (Eurostat, 2011a), although population growth is slow (total and working age population) (Rees et al., 2012). Ageing of the population is a significant trend in Europe, as in all high income populations. This will have both economic and social implications, with many regions experiencing a decline in the labour force (Rees et al., 2012). Since AR4, economic growth has slowed or become negative in many countries, leading to a reduction in social protection measures and increased unemployment (Eurostat, 2011b). The longer term implications of the financial crisis in Europe are unclear, although it may lead to a modification of the economic outlook and affect future social protection policies with implications for adaptation. Europe is one of the world s largest and most productive suppliers of food and fibre (Easterling et al., 2007) and agriculture is the most important European land use by area (45% of the total area) (Rounsevell et al., 2006). After 1945, an unprecedented increase in agricultural productivity occurred, but also declines in agricultural land use areas. This intensification had several negative impacts on the ecological properties of agricultural systems, such as carbon sequestration, nutrient cycling, soil structure and functioning, water purification and pollination. Pollution from agriculture has led to eutrophication and declines in water quality in some areas (ELME, 2007). Most scenario studies suggest that agricultural land areas will continue to decrease in the future (see also (Busch, 2006) for a discussion). Agriculture accounts for 24 % of total national freshwater abstraction in Europe and more than 80 % in some southern European countries (EEA, 2009). Economic restructuring in some eastern European countries has led to a decrease in water abstraction for irrigation, suggesting the potential for future increases in irrigated agriculture and water use efficiency (EEA, 2009). Forest in Europe covers approximately 35% of the land area (Eurostat, 2009). The majority of forests now grow faster than in the early 20th century due to advances in forest management practices, genetic improvement and in central Europe, the cessation of site-degrading practices such as litter collection for fuel. Increasing temperatures and CO2 concentrations, nitrogen deposition, and the reduction of air pollution (SO2) have also had a positive effect on forest growth. Scenario studies suggest that forested areas will increase in Europe in the future on land formerly used for agriculture (Rounsevell et al., 2006). Soil degradation is already intense in parts of the Mediterranean and central-eastern Europe and, together with prolonged drought periods and fires, is already contributing to an increased risk of desertification. Projected risks for future desertification are the highest in these areas (EEA, 2012). Urban development is projected to increase all over Europe (Reginster and Rounsevell, 2006), but especially rapidly in Eastern Europe, with the magnitude of these increases depending on population growth, economic growth and land use planning policy. Although changes in urban land use will be relatively small in area terms, urban development has major impacts locally on environmental quality. Outdoor air quality has, however, been improving (ELME, 2007). Peri-urbanisation is an increasing trend in which residents move out of cities to locations with a rural character, but retain a functional link to cities by commuting to work (Reginster and Rounsevell, 2006)(Rounsevell and Reay, 2009). Several European scenario studies have been undertaken to describe European future trends with respect to: socio- economic development (Mooij de and Tang, 2003), land use change (Letourneau et al., 2012; Verburg et al., 2010)(Haines-Young et al., 2012), land use and biodiversity (Spangenberg et al., 2011), crop production (Hermans et al., 2010), demographic change (Davoudi et al., 2010), economic development (Dammers, 2010) and European policy (Helming et al., 2011)(Lennert and Robert, 2010). Many of these scenarios also account for the effects of Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 future climate change (see (Rounsevell and Metzger, 2010) for a review). Long term projections (to the end of the century) are described under the new Shared Socio-economic Pathway scenarios (SSPs) (Kriegler et al., 2010). Detailed country and regional scale socio-economic scenarios have also been produced for the Netherlands (WLO, 2006), the UK (UK National Ecosystem Assessment, 2011) and Scotland (Harrison et al., 2013). The probabilistic representation of socio-economic futures has also been developed for agricultural land use change (Hardacre et al., 2012). There is little evidence to suggest, however, that probabilistic futures or scenarios more generally are being used in policy making (Bryson et al., 2010). 23.2.2. Observed and Projected Climate Change 23.2.2.1. Observed Climate Change The average temperature in Europe has continued to increase with regionally and seasonally different rates of warming, being greatest in high latitudes in Northern Europe (AR5 WG2 Chapter 28). Since the 1980s, warming has been strongest over Scandinavia, especially in winter, whereas the Iberian Peninsula warmed mostly in summer (EEA, 2012; Haylock et al., 2008). The decadal average temperature over land area for 2002-2011 is 1.3°C+/- 0.11°C above the 1850-1899 average, based on HadCRUT3 (Brohan et al., 2006), MLOST (Smith et al., 2008) and GISS Temp (Hansen et al., 2010). See AR5 WG1 Section 2.4 for a discussion of data and uncertainties and AR5 WG2 Chapter 21for observed regional climate change. Since 1950, high-temperature extremes (hot days, tropical nights, and heat waves) have become more frequent, while low-temperature extremes (cold spells, frost days) have become less frequent (AR5 WG1 Chapter 2.6, SREX- 3)(EEA, 2012). The recent cold winters in Northern and Atlantic Europe reflect the high natural variability in the region (Peterson et al., 2012)(AR5 WG1 section 2.7), and do not contradict the general warming trend. In Eastern Europe, including the European part of Russia, summer 2010 was exceptionally hot, with an amplitude and spatial extent that exceeded the previous 2003 heat wave (Barriopedro et al., 2011). Table 23-1 describes the impacts of major extreme events in Europe in the last decade. Since 1950, annual precipitation has increased in Northern Europe (up to +70 mm/decade) based on Haylock et al. (2008), and decreased in parts of Southern Europe (EEA, 2012). Winter snow cover extent has a high inter-annual variability and a non-significant negative trend over the period 1967-2007 (Henderson and Leathers, 2010). Regional observed changes in temperature and precipitation extremes are also described in Table 3-2 of SREX and in Berg et al. (2013). Mean wind speeds have declined over Europe over recent decades (Vautard et al., 2010) with low confidence due to problematic anemometer data and climate variability (SREX Section 3.3). Bett et al (2013) did not find any trend in windspeed using the Twentieth Century Reanalysis. Europe is marked by increasing mean sea level with regional variations, except in the northern Baltic Sea where the relative sea level decreased due to vertical crustal motion (Albrecht et al., 2011; EEA, 2012; Haigh et al., 2010; Menendez and WoodWorth, 2010). Extreme sea levels have increased due to mean sea level rise (medium confidence, SREX Section 3.5, Haigh et al., 2010; Menendez and WoodWorth, 2010). Variability in waves is related to internal climate variability rather than climate trends (SREX Section 3.5, Charles et al., 2012). 23.2.2.2. Projected Climate Changes For Europe, sub-regional information from global (AR5 WG1 Chapter 14.8.6; AR5 WG1 Annex 1; AR5 WG2 Chapter 21 supplement) and regional high resolution climate model output (AR5 WG1 Chapter 14.8.6; WG2 Chapter 21, 23) provide more knowledge about the range of possible future climates under the SRES and RCP emission scenarios. Within the recognized limitations of climate projections (AR5 WG1 Chapter 9; WG2 Chapter 21), new research on inter-model comparisons has provided a more robust range of future climates to assess future impacts. Since AR4, climate impact assessments are more likely to use a range for the projected changes in temperature and rainfall. Access to comprehensive and detailed sets of climate projections for decision making exist in Europe (SREX Section 3.2.1, (Mitchell et al., 2004)(Fronzek et al., 2012; Jacob et al., 2013). Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climate models show significant agreement for all emission scenarios in warming (magnitude and rate) all over Europe, with strongest warming projected in Southern Europe in summer, and in Northern Europe in winter (Kjellström et al., 2011)(Goodess et al., 2009). Even under an average global temperature increase limited to 2°C compared to pre-industrial times, the climate of Europe is simulated to depart significantly in the next decades from today s climate (Jacob and Podzun, 2010);(Van der Linden and Mitchell, 2009). Precipitation signals vary regionally and seasonally. Trends are less clear in Continental Europe, with agreement in increase in Northern Europe and decrease in Southern Europe (medium confidence) (Kjellström et al., 2011). Precipitation is projected to decrease in the summer months up to Southern Sweden and increase in winter (Schmidli et al., 2007) with more rain than snow in mountainous regions (Steger et al., 2013). In Northern Europe, a decrease of long term mean snow pack (although snow-rich winters will remain) towards the end of the century (Räisänen and Eklund, 2012) is projected. There is lack of information about past and future changes in hail occurrence in Europe. Changes in future circulation patterns (Kreienkamp et al., 2010; Ulbrich et al., 2009) and mean wind speed trends are uncertain in sign (Kjellström et al., 2011)(McInnes et al., 2011). Regional coupled simulations over the Mediterranean region provide a more realistic characterization of impact parameters (e.g. snow cover, aridity index, river discharge), which were not revealed by CMIP3 global simulations (Dell'Aquila et al., 2012). For 2081-2100 compared to 1986-2005, projected global mean sea level rises (metres) are in the range 0.29-0.55 for RCP2.6, 0.36-0.63 for RCP4.5, 0.37-0.64 for RCP6.0 and 0.48-0.82 for RCP8.5 (medium confidence, AR5 WG3 Chapter 5). There is a low confidence on projected regional changes (Slangen et al., 2012)(AR5 WG1 13.6). Low probability/high impact estimates of extreme mean sea-level rise projections derived from the A1FI SRES scenario for the Netherlands (Katsman et al., 2011) indicate that the mean sea-level could rise globally between 0.55 and 1.15 m, and locally (the Netherlands) by 0.40 to 1.05 m, by 2100. Extreme (very unlikely) scenarios for the UK vary from 0.9 to 1.9 m by 2100 (Lowe et al., 2009). 23.2.2.3. Projected Changes in Climate Extremes There will be a marked increase in extremes in Europe, in particular, in heat waves, droughts and heavy precipitation events (Beniston et al., 2007)(Lenderink and Van Meijgaard, 2008) and AR5 WG2 Chapter 21 Supplement. There is a general high confidence concerning changes in temperature extremes (toward increased number of warm days, warm nights and heat waves, SREX Table 3-3). Figure 23-2 (upper panels) shows projected changes in the mean number of heat waves in May to September for 2071-2100 compared to 1971-2000 for RCP4.5 and RCP8.5 with large differences depending on the emission scenario. The increase in likelihood of some individual events due to anthropogenic change has been quantified for the 2003 heat wave (Schär and Jendritzky, 2004), the warm winter of 2006/2007 and warm spring of 2007 (Beniston, 2007). Changes in extreme precipitation depend on the region, with a high confidence of increased extreme precipitation in Northern Europe (all seasons) and Continental Europe (except summer). Future projections are regionally and seasonally different in Southern Europe (SREX Table 3-3). Figure 23-2 (middle panels) shows projected seasonal changes of heavy precipitation events for 2071-2100 compared to 1971-2000 for RCP4.5 and RCP8.5. [INSERT FIGURE 23-2 HERE Figure 23-2: First row: Projected changes in the mean number of heat waves occurring in the months May to September for the period 2071-2100 compared to 1971-2000 (number per 30 years). Heat waves are defined as periods of more than 5 consecutive days with daily maximum temperature exceeding the mean maximum temperature of the May to September season of the control period (1971-2000) by at least 5°C. Second and third rows: Projected seasonal changes in heavy precipitation defined as the 95th percentile of daily precipitation (only days with precipitation > 1mm/day are considered) for the period 2071-2100 compared to 1971-2000 (in %) in the months December to January (DJF) and June to August (JJA). Fourth row: Projected changes in the 95th percentile of the length of dry spells for the period 2071-2100 compared to 1971-2000 (in days). Dry spells are defined as Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 periods of at least 5 consecutive days with daily precipitation below 1mm. Hatched areas indicate regions with robust (at least 66% of models agree in the sign of change) and/or statistical significant change (significant on a 95% confidence level using Mann-Whitney-U test). For the eastern parts of Black Sea, Eastern Anatolia and Southeast Anatolia (Turkey), no regional climate model projections are available. Changes represent the mean over 8 (RCP4.5, left side) and 9 (RCP8.5, right side) regional model simulations compiled within the EURO-CORDEX initiative. Adapted from Jacob et al. (2013).] Projected changes of spatially averaged indices over the European sub-regions (Figure 23-1) are described in the supplemental information (Table SM23-2). In winter, small increases in extreme wind speed are projected for Central and Northern Europe [medium confidence] (AR5 WG2 21.3.3.1.6; SREX Figure 3-8) (Beniston et al., 2007; Haugen and Iversen, 2008; Rauthe et al., 2010; Rockel and Woth, 2007; Schwierz et al., 2010), connected to changes in storm tracks [medium confidence] (Pinto et al., 2007a; Pinto et al., 2007b)(Donat et al., 2010)(Pinto et al., 2010). Other parts of Europe and seasons are less clear in sign with a small decreasing trend in southern Europe [low confidence] (Donat et al., 2011; McInnes et al., 2011). Extreme sea level events will increase (high confidence, AR5 WG1 13.7, SREX 3.5.3), mainly dominated by the global mean sea level increase. Storm surges are expected to vary along the European coasts. Significant increases are projected in the eastern North Sea (increase of 6-8% of the 99th percentile of the storm surge residual, 2071-2100 compared to 1961-1990, based on the B2, A1B and A2 SRES scenarios) (Debernard and Ryed, 2008) and west of UK and Ireland (Debernard and Ryed, 2008)(Wang et al., 2008), except South of Ireland (Wang et al., 2008). There is a medium agreement for the South of North Sea and Dutch coast where trends vary from increasing (Debernard and Ryed, 2008) to stable (Sterl et al., 2009). There is a low agreement on the trends in storm surge in the Adriatic sea (Jorda et al., 2012; Lionello et al., 2012; Troccoli et al., 2012b)(Planton et al., 2011). 23.2.3. Observed and Projected Trends in the Riverflow and Drought Streamflows have decreased in the south and east of Europe and increased in Northern Europe (Stahl et al., 2010)(Wilson et al., 2010) (AR5 WG2 3.2.3). In general, few changes in flood trends can be attributed to climate change, partly due to the lack of sufficiently long records (Kundzewicz et al., 2013). European mean and peak discharges are highly variable (Bouwer et al., 2008); for instance in France, upward trends in low flows were observed over 1948-1988 and downward trends over 1968-2008 (Giuntoli et al., 2013). Alpine glacier retreat during the last two decades caused a 13% increase in glacier contribution to August runoff of the four main rivers originating in the Alps, compared to the long-term average (Huss, 2011). Increases in extreme river discharge (peak flows) over the past 30-50 years have been observed in parts of Germany (Petrow et al., 2009)(Petrow et al., 2007), the Meuse river basin (Tu et al., 2005), parts of Central Europe (Villarini et al., 2011), Russia (Semenov, 2011), and Northeastern France (Renard et al., 2008). Decreases in extreme river discharge have been observed in the Czech Republic (Yiou et al., 2006), and no change observed in Switzerland (Schmocker-Fackel and Naef, 2010), Germany (Bormann et al., 2011), and the Nordic countries (Wilson et al., 2010). River regulation possibly partly masks increasing peak flows in the Rhine (Vorogushyn et al., 2012). One study (Pall et al., 2011) suggested that the UK 2000 flood was partly due to anthropogenic forcing, although another showed a weaker effect (Kay et al., 2011). Climate change is projected to affect the hydrology of river basins (SREX Chapter 3; AR5 WG2 Chapter 4). The occurrence of current 100-year return period discharges is projected to increase in Continental Europe, but decrease in some parts of Northern and Southern Europe by 2100 (Dankers and Feyen, 2008)(Rojas et al., 2012). In contrast, studies for individual catchments indicate increases in extreme discharges, to varying degrees, in Finland (Veijalainen et al., 2010), Denmark (Thodsen, 2007), Ireland (Wang et al., 2006)(Steele-Dunne et al., 2008)(Bastola et al., 2011), the Rhine basin (Görgen et al., 2010; Te Linde et al., 2010a), Meuse basin (Leander et al., 2008)(Ward et al., 2011), the Danube basin (Dankers et al., 2007), and France (Chauveau et al., 2013; Quintana-Segui et al., 2011). Although snowmelt floods may decrease, increased autumn and winter rainfall could lead to higher peak discharges in northern Europe (Lawrence and Hisdal, 2011). Declines in low flows are projected for the UK Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (Christierson et al., 2012), Turkey (Fujihara et al., 2008), France (Chauveau et al., 2013), and rivers fed by Alpine glaciers (Huss, 2011). The analysis of trends in droughts is made complex by the different categories or definitions of drought (meteorological, agricultural, and hydrological) and the lack of long-term observational data (SREX Box 3-3). Southern Europe shows trends towards more intense and longer meteorological droughts, but they are still inconsistent (Sousa et al., 2011). Drought trends in all other sub-regions are not statistically significant (SREX 3.5.1). Regional and global climate simulations project (medium confidence) an increase in duration and intensity of droughts in central and southern Europe and the Mediterranean up until the UK for different definitions of drought (Feyen and Dankers, 2009; Gao and Giorgi, 2008; Vidal and Wade, 2009)(Koutroulis et al., 2010; Tsanis et al., 2011) (AR5 WG2 Chapter 21). Even in regions where summer precipitation is expected to increase, soil moisture and hydrological droughts may become more severe due to increasing evapotranspiration (Wong et al., 2011). Projected changes in the length of meteorological dry spells show that the increase is large in Southern Europe (Figure 23-2 fourth row). 23.3. Implications of Climate Change for Production Systems and Physical Infrastructure 23.3.1. Settlements 23.3.1.1. Coastal Flooding As the risk of extreme sea level events increases with climate change [23.2.3, AR5 WG2 Chapter 5], coastal flood risk will remain a key challenge for several European cities, port facilities and other infrastructure (Nicholls et al., 2008)(Hallegatte et al., 2008)(Hallegatte et al., 2011). With no adaptation, coastal flooding in the 2080s is projected to affect an additional 775,000 and 5.5 million people per year in the EU27 (B2 and A2 scenarios) (Ciscar et al., 2011). The Atlantic, Northern and Southern European regions are projected to be most affected. Direct costs from sea level rise in the EU27 without adaptation could reach 17 billion Euros per year by 2100 (Hinkel et al., 2010), with indirect costs also estimated for land-locked countries (Bosello et al., 2012). Countries with high absolute damage costs include the Netherlands, Germany, France, Belgium, Denmark, Spain and Italy (Hinkel et al., 2010). Upgrading coastal defences would substantially reduce impacts and damage costs (Hinkel et al., 2010). However, the amount of assets and populations that need to protected by coastal defences is increasing, thus, the magnitude of losses when floods do occur will also increase in the futre (Hallegatte et al. 2013), entailing the need to prepare for very large flood disasters in the future. An increase in future flood losses due to climate change have been estimated for Copenhagen (Hallegatte et al., 2011), the UK coast (Mokrech et al., 2008)(Purvis et al., 2008)(Dawson et al., 2011), the North Sea coast (Gaslikova et al., 2011), cities including Amsterdam and Rotterdam (Hanson et al., 2011), and the Netherlands (Aerts et al., 2008). A 1m sea-level rise in Turkey could affect 3 million additional people and put 12 billion USD capital value at risk, with around 20 billion USD adaptation costs (10% of GNP) (Karaca and Nicholls, 2008). In Poland, up to 240,000 people would be affected by increasing flood risk on the Baltic coast (Pruszak and Zawadzka, 2008). The increasing cost of insurance and unwillingness of investors to place assets in affected areas is a potential growth impediment to coastal and island economies (Day et al., 2008). 23.3.1.2. River and Pluvial Flooding Recent major flood events in Europe include the 2007 floods in the UK (Table 23-1) (Chatterton et al., 2010) and the 2013 floods in Germany. The observed increase in river flood events and damages in Europe is well documented (see AR5 WG2 18.4.2.1), however, the main cause is increased exposure of persons and property in flood risk areas (Barredo, 2009). Since AR4, new studies provide a wider range of estimates of future economic losses from river flooding attributable to climate change, depending on the modelling approach and climate scenario (Bubeck et al., 2011). Studies now also quantify risk under changes in population and economic growth, generally indicating this contribution to be about equal or larger than climate change per se (Feyen et al., 2009; Maaskant et al., 2009; Rojas Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 et al., 2013)(Bouwer et al., 2010)(Te Linde et al., 2011). Some regions may see increasing risks, but others may see decreases or little to no change (Bubeck et al., 2011)(ABI, 2009)(Feyen et al., 2009)(Lugeri et al., 2010)(Mechler et al., 2010)(Feyen et al., 2012)(Lung et al., 2012). In the EU15, river flooding could affect 250,000-400,000 additional people by the 2080s (SRES A2 and B2 scenarios) and more than doubling annual average damages, with Central and Northern Europe and the UK most affected (Ciscar, 2009)(Ciscar et al., 2011). When economic growth is included, economic flood losses in Europe could increase 17-fold under the A1B climate scenario (Rojas et al., 2013). Few studies have estimated future damages from inundation in response to an increase in intense rainfall (Hoes, 2006; Willems et al., 2012). Processes that influence flash flood risk include increasing exposure from urban expansion, and forest fires that lead to erosion and increased surface runoff (Lasda et al., 2010). Some studies have costed adaptation measures but these may only partly offset anticipated impacts (Zhou et al., 2012). [INSERT Table 23-1 HERE Table 23-1: Impacts of climate extremes in the last decade in Europe.] 23.3.1.3. Windstorms Several studies project an overall increase storm hazard in northwest Europe [23.2.2.3] and in economic and insured losses [AR5 WG2 Chapter 17.7.3], but natural variations in frequencies are large. There is no evidence that the observed increase in European storm losses is due to anthropogenic climate change (Barredo, 2010). There is a lack of information for other storm types, such as tornadoes and thunderstorms. 23.3.1.4.Mass Movements and Avalanches In the European Alps, the frequency of rock avalanches and large rock slides has apparently increased over the period 1900-2007 (Fischer et al., 2012). The frequency of landslides may also have increased in some locations (Lopez Saez et al., 2013). Mass movements are projected to become more frequent with climate change (Huggel et al., 2010; Stoffel and Huggel, 2012), although several studies indicate a more complex or stabilising response of mass movements to climate change (Dixon and Brook, 2007; Huggel et al., 2012; Jomelli et al., 2007; Jomelli et al., 2009; Melchiorre and Frattini, 2012). Some land-use practices have led to conditions favourable to increased landslide risk, despite climate trends that would result in a decrease of landslide frequency, as reported in Calabria (Polemio and Petrucci, 2010) and in the Apennines (Wasowski et al., 2010). Snow avalanche frequency changes in Europe are dominated by climate variability; studies based on avalanche observations (Eckert et al., 2010) or favourable meteorological conditions (Castebrunet et al., 2012; Teich et al., 2012) show contrasting variations, depending on the region, elevation, season and orientation. 23.3.2. Built Environment Built infrastructure in Europe is vulnerable to extreme weather events, including overheating of buildings (houses, hospitals, schools) during hot weather (Crump et al., 2009; DCLG, 2012). Buildings that were originally designed for certain thermal conditions will need to function in warmer climates in the future (WHO, 2008). Climate change in Europe is expected to increase cooling energy demand (23.3.4) (Dolinar et al., 2010), with implications for mitigation and adaptation policies (23.8.1). A range of adaptive strategies for buildings are available, including effective thermal mass and solar shading (ARUP, 2008). Climate change may also increase the frequency and intensity of drought-induced soil subsidence and associated damage to dwellings (Corti et al., 2009). With respect to the outdoor built environment, there is limited evidence regarding the potential for differential rates of radiatively-forced climate change in urban compared to rural areas (McCarthy et al., 2010). Climate change may exacerbate London's nocturnal urban heat island (UHI) (Wilby, 2008), however, the response of different cities may vary. For example, a study of Paris (Lemonsu et al., 2013) indicated a future reduction in strong urban heat island Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 events when increased soil dryness was taken into effect. Modification of the built environment, via enhanced urban greening, for example, can reduce temperatures in urban areas, with co-benefits for health and wellbeing (23.7.4, 23.8.1). 23.3.3. Transport Systematic and detailed knowledge on climate change impacts on transport in Europe remains limited (Koetse and Rietveld, 2009). On road transport, in line with AR4, more frequent but less severe collisions due to reduced speed are expected in case of increased precipitation (Brijs et al., 2008)(Kilpeläinen and Summala, 2007). However, lower traffic speed may cause welfare losses due to additional time spent driving (Sabir et al., 2010). Severe snow and ice-related accidents will also decrease, but the effect of fewer frost days on total accidents is unclear (Andersson and Chapman, 2011a)(Andersson and Chapman, 2011b). Severe accidents caused by extreme weather are projected to decrease by 63-70% in 2040-2070 compared to 2007 as a result of modified climate and expected developments in vehicle technology and emergency systems (Nokkala et al., 2012). For rail, consistent with AR4, increased buckling in summer, as occurred in 2003 in the UK, is expected to increase the average annual cost of heat-related delays in some regions, while the opposite is expected for ice and snow- related delays (Dobney et al., 2010; Lindgren et al., 2009; Palin et al., 2013). Effects from extreme precipitation, as well as the net overall regional impact of climate change remain unclear. Efficient adaptation comprises proper maintenance of track and track bed. Regarding inland waterways, the case of Rhine shows that for 1-2 oC increases by 2050 more frequent high water levels are expected in winter, while after 2050 days with low water levels in summer will also increase (Jonkeren et al., 2011)(Te Linde et al., 2011)(Te Linde, 2007)(Hurkmans et al., 2010). Low water levels will reduce the load factor of inland ships and consequently increase transport prices, as in the Rhine and Moselle in 2003 (Jonkeren, 2009)(Jonkeren et al., 2007). Adaptation includes modal shifts, increase navigational hours per day under low water levels, and infrastructure modifications (e.g. canalization of river parts) (Jonkeren et al., 2011; Krekt et al., 2011). For long range ocean routes, the economic attractiveness of the Northwest Passage and the Northern Sea Route depends also on passage fees, bunker prices and cost of alternative sea routes (Verny and Grigentin, 2009)(Liu and Kronbak, 2010)(Lasserre and Pelletier, 2011). Regarding air transport, for Heathrow airport in the UK, future temperature and wind changes were estimated to cause a small net annual increase but much larger seasonal changes on the occurrence of delays (Pejovic et al., 2009). 23.3.4. Energy Production, Transmission, and Use On wind energy, no significant changes are expected before 2050, at least in Northern Europe (Pryor and Schoof, 2010)(Pryor and Barthelmie, 2010)(Seljom et al., 2011)(Barstad et al., 2012; Hueging et al., 2013). After 2050, in line with AR4, the wind energy potential in Northern, Continental and most of Atlantic Europe may increase during winter and decrease in summer (Harrison et al., 2008; Hueging et al., 2013)(Nolan et al., 2012; Rockel and Woth, 2007). For Southern Europe, a decrease in both seasons is expected, except for the Aegean Sea and Adriatic coast where a significant increase during summer is possible (Bloom et al., 2008; Hueging et al., 2013; Najac et al., 2011; Pa¹ièko et al., 2012). For hydropower, electricity production in Scandinavia is expected to increase by 5-14% during 2071-2100 compared to historic or present levels (Golombek et al., 2012) (Haddeland et al., 2011); for 2021-2050, increases by 1-20% were estimated (Haddeland et al., 2011)(Hamududu and Killingtveit, 2012; Seljom et al., 2011). In Continental, and part of Alpine Europe, reductions in electricity production by 6-36% were estimated (Schaefli et al., 2007) (Paiva et al., 2011; Pa¹ièko et al., 2012)(Hendrickx and Sauquet, 2013; Stanzel and Nachtnebel, 2010). For Southern Europe, production is expected to decrease by 5-15% in 2050 compared to 2005 (Bangash et al., 2013; Hamududu and Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Killingtveit, 2012). Adaptation consists in improved water management, including pump storage if appropriate (Schaefli et al., 2007)(García-Ruiz et al., 2011). Biofuel production is discussed in section 23.4.5. There are few studies of impacts on solar energy production. Crook et al. (2011) estimated an increase of the energy output from photovoltaic panels and especially from concentrated solar power plants in most of Europe under the A1B scenario. On thermal power, in line with AR4, van Vliet et al. (2012) estimated a 6-19% decrease of the summer average usable capacity of power plants by 2031 2060 compared to 1971-2000, while smaller decreases have been also estimated (Linnerud et al., 2011)(Förster and Lilliestam, 2010). Closed-cooling circuits are efficient adaptation choices for new plants (Koch and Vögele, 2009). In power transmission, increasing lightning and decreasing snow- sleet-and blizzard faults for 2050-2080 were estimated for the UK (McColl et al., 2012). By considering both heating and cooling, under a +3.7 oC scenario by 2100 a decrease of total annual energy demand in Europe as a whole during 2000-2100 was estimated (Isaac and van Vuuren, 2009). Seasonal changes will be prominent, especially for electricity (see Figure 23-3), with summer peaks arising also in countries with moderate summer temperatures (Hekkenberg et al., 2009). Heating degree days are expected to decrease by 11-20% between 2000 and 2050 due solely to climate change (Isaac and van Vuuren, 2009). For cooling, very large percentage increases up to 2050 are estimated by the same authors for most of Europe as the current penetration of cooling devices is low; then, increases by 74-118% in 2100 (depending on the region) from 2050 are expected under the combined effect of climatic and non-climatic drivers. In Southern Europe, cooling degree days by 2060 will increase, while heating degree days will decrease but with substantial spatial variations (Giannakopoulos et al., 2009). Consequently, net annual electricity generation cost will increase in most of the Mediterranean and decrease in the rest of Europe (Eskeland and Mideksa, 2010)(Mirasgedis et al., 2007)(Pilli-Sihlova et al., 2010; Zachariadis, 2010). Future building stock changes and retrofit rates are critical for impact assessment and adaptation (Olonscheck et al., 2011). Energy efficient buildings and cooling systems, and demand-side management are effective adaptation options (Artmann et al., 2008; Breesch and Janssens, 2010; Chow and Levermore, 2010; Day et al., 2009; Jenkins et al., 2008). [INSERT FIGURE 23-3 HERE Figure 23-3: Percentage change in electricity demand in Greece attributable to climate change, under a range of climate scenarios and economic assumptions. Source: Mirasgedis et al., 2007.] 23.3.5. Industry and Manufacturing Research on the potential effects of climate change in industry is limited. Modifications in future consumption of food and beverage products have been estimated on the basis of current sensitivity to seasonal temperature (Mirasgedis et al., 2013). Higher temperatures may favour the growth of food borne pathogens or contaminants (Jacxsens et al., 2010; Popov Janevska et al., 2010) (see also 23.5.1). The quality of some products, such as wine (23.4.1, Box 23-2), is also likely to be affected. In other sectors, the cumulative cost of direct climate change impacts in the Greek mining sector for 2021-2050 has been estimated at 0.245 billion Euros, in 2010 prices (Damigos, 2012). Adaptation to buildings or work practices are likely to be needed in order to maintain labour productivity during hot weather (Kjellstrom et al., 2009)(11.6.2.2). 23.3.6. Tourism In line with AR4, the climate for general tourist activities especially after 2070 is expected to improve significantly during summer and less during autumn and spring in northern Continental Europe, Finland, southern Scandinavia and southern England (Amelung and Moreno, 2012)(Amelung et al., 2007)(Nicholls and Amelung, 2008). For the Mediterranean, climatic conditions for light outdoor tourist activities are expected to deteriorate in summer mainly after 2050, but improve during spring and autumn (Amelung and Moreno, 2009)(Hein et al., 2009)(Perch-Nielsen et al., 2010)(Amelung et al., 2007)(Giannakopoulos et al., 2011). Others concluded that before 2030 (or even 2060) Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 this region as a whole will not become too hot for beach or urban tourism (Moreno and Amelung, 2009)(Rutty and Scott, 2010), while surveys showed that beach tourists are deterred mostly by rain (De Freitas et al., 2008; Moreno, 2010). Thus, from 2050, domestic tourism and tourist arrivals at locations in Northern and parts of Continental Europe may be enhanced at the expense of Southern locations (Amelung and Moreno, 2012; Bujosa and Roselló, 2012; Hamilton and Tol, 2007; Hein et al., 2009). The age of tourists, the climate in their home country, local economic and environmental conditions (e.g. water stress, tourist development) are also critical (Hamilton and Tol, 2007)(Moreno and Amelung, 2009; Perch-Nielsen et al., 2010)(Eugenio-Martin and Campos-Soria, 2010; Lyons et al., 2009)(Rico- Amoros et al., 2009). Tourism in mountainous areas may benefit from improved climatic conditions in summer (Endler et al., 2010; Endler and Matzarakis, 2011b; Perch-Nielsen et al., 2010; Serquet and Rebetez, 2011). However, in agreement with AR4, natural snow reliability and thus ski season length will be adversely affected, especially where artificial snowmaking is limited (OECD, 2007; Steiger, 2011)(Moen and Fredman, 2007). Low-lying areas will be the most vulnerable (Endler et al., 2010; Endler and Matzarakis, 2011a; Serquet and Rebetez, 2011; Steiger, 2011; Uhlmann et al., 2009). Tourist response to marginal snow conditions remains largely unknown, while changes in weather extremes may also be critical (Tervo, 2008). Up to 2050, demographic changes (e.g. population declines in source countries, ageing populations) may have a higher impact than climate change (Steiger, 2012). Artificial snowmaking has physical and economic limitations, especially in small sized and low-altitude ski stations (Sauter et al., 2010; Steiger and Mayer, 2008; Steiger, 2010; Steiger, 2011), and increases water and energy consumption. Shifts to higher altitudes, operational/ technical measures and year-round tourist activities may not fully compensate for adverse impacts. 23.3.7. Insurance and Banking Insurance and banking face problems related to accurate pricing of risks, shortage of capital after large loss events, and by an increasing burden of losses that can affect markets and insurability, within but also outside the European region (Botzen et al., 2010a; Botzen et al., 2010b; CEA, 2007)(AR5 WG2 Section 10.7). However, risk transfer including insurance also holds potential for adaptation by providing incentives to reduce losses (Botzen and van den Bergh, 2008; CEA, 2009)(Herweijer et al., 2009). Banking is potentially affected through physical impacts on assets and investments, as well as through regulation and/or mitigation actions by changing demands regarding sustainability of investments and lending portfolios. Few banks have adopted climate strategies that also address adaptation (Furrer et al., 2009)(Cogan, 2008). Windstorm losses are well covered in Europe by building and motor policies, and thus create a large exposure to the insurance sector. Flood losses in the UK in 2000, 2007 and 2009 have put the insurance market under further pressure, with increasing need for the government to reduce risk (Ward et al., 2008)(Lamond et al., 2009). Other risks of concern to the European insurance industry is building subsidence related to drought (Corti et al., 2009), and hail damage to buildings and agriculture (Kunz et al., 2009; Botzen et al., 2010b; GIA, 2011). The financial sector can adapt through adjustment of premiums, restricting or reduction of coverage, further risk spreading, and importantly incentivising risk reduction (Botzen et al., 2010a; Clemo, 2008)(Crichton, 2007)(Crichton, 2006)(Wamsler and Lawson, 2011)(Surminski and Philp, 2010). Public attitudes in Scotland and the Netherlands would support insurance of private property and public infrastructure damages in the case of increasing flood risk (Botzen et al., 2009)(Glenk and Fisher, 2010). Government intervention is however often needed to provide compensation and back-stopping in the event of major losses (Aakre and Rübbelke, 2010; Aakre et al., 2010). Hochrainer et al. (2010) analysed the performance of the EU Solidarity Fund that supports European governments in large events, and argue there is a need to increase its focus on risk reduction. Current insurance approaches present in Europe are likely to remain, as they are tailored to local situations and preferences (Schwarze et al., 2011). Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 23.4. Implications of Climate Change for Agriculture, Fisheries, Forestry, and Bioenergy Production 23.4.1. Plant (Food) Production In AR4, Alcamo et al. (2007) reported that crop suitability is likely to change throughout Europe. During the 2003 and 2010 summer heat waves, grain-harvest losses reached 20 and 30% in affected regions of Europe and Russia, respectively (Barriopedro et al., 2011; Ciais et al., 2005) (Table 23-1). Cereals production fell on average by 40% in the Iberian Peninsula during the intense 2004/2005 drought (EEA, 2010a). Climate-induced variability in wheat production has increased in recent decades in Southern and Central Europe (Brisson et al., 2010)(Hawkins et al., 2013)(Ladanyi, 2008), but no consistent reduction has been recorded in the northernmost areas of Europe (Peltonen- sainio et al., 2010). Country-scale rainfed cereals yields are below agro-climatic potentials (Supit et al., 2010) and wheat yield increases have levelled off in several countries over 1961-2009 (Olesen et al., 2011). High temperatures and droughts during grain filling has contributed to the lack of yield increase of winter wheat in France despite improvements in crop breeding (Brisson et al., 2010; Kristensen et al., 2011). In contrast, in eastern Scotland, warming has favoured an increase in potato yields since 1960 (Gregory and Marshall, 2012). In north-east Spain, grape yield was reduced by an increased water deficit in the reproductive stage since the 1960s (Camps and Ramos, 2012). Insight into the potential effect of climate change on crops requires the combination of a wide range of emission scenarios, global circulation models (GCM) and impact studies (Trnka et al., 2007)(Soussana et al., 2010). In the EU27, a 2.5 °C regional temperature increase in the 2080s under the B2 scenario could lead to small changes (on average +3%) in crop yields, whereas a 5.4 °C regional warming under the A2 scenario could reduce mean yields by 10% according to a study based on regional climate models (Ciscar et al., 2011). An initial benefit from the increasing CO2 concentration for rainfed crop yields would contrast by the end of the century with yield declines in most European subregions, although wheat yield could increase under the A2 scenario (Supit et al., 2012, three GCMs, B1, A2 scenarios). Disease-limited yields of rain fed wheat and maize in the 2030s does not show consistent trends across two GCMs (Donatelli et al., 2012). For a global temperature increase of 5° C, agroclimatic indices show an increasing frequency of extremely unfavourable years in European cropping areas (Trnka et al., 2011). Under the A2 and B2 scenarios, crop production shortfalls, defined as years with production below 50% of its average climate normal production would double by 2020 and triple by 2070 as compared to a current frequency of 1-3 years per decade in the currently most productive southern European regions of Russia (Alcamo et al., 2007). The regional distribution of climate change impacts on agricultural production is likely to vary widely (Iglesias et al., 2012) (Donatelli et al., 2012) (Figure 23-4). Southern Europe would experience the largest yield losses (-25 % by 2080 under a 5.4 °C warming, (Ciscar et al., 2011) with increased risks of rain fed summer crop failure (Bindi and Olesen, 2011)(Ferrara et al., 2010)(Ruiz-Ramos et al., 2011). Warmer and drier conditions by 2050 (Trnka et al., 2010; Trnka et al., 2011) would cause moderate declines in crop yields in Central Europe regions (Ciscar et al., 2011). In Western Europe, increased heat stress around flowering could cause considerable yield losses in wheat (Semenov, 2009). For Northern Europe, there is diverging evidence concerning future impacts. Positive yield changes combined with the expansion of climatically suitable areas could lead to crop production increases (between 2.5 and 5.4°C regional warming) (Bindi and Olesen, 2011)(Bindi and Olesen, 2011). However, increased climatic variability would limit winter crops expansion (Peltonen-Sainio et al., 2010) and cause at high latitudes high risk of marked cereal yield loss (Rötter et al., 2011). Spring crops from tropical origin like maize for silage could become cultivated in Finland by the end of the century (Peltonen-Sainio et al., 2009). Cereal yield reduction from ozone (Fuhrer, 2009) could reach 6 and 10 % in 2030 for the European Union with the B1 and A2 scenarios, respectively (Avnery et al., 2011a; Avnery et al., 2011b). Because of limited land availability and soil fertility outside of Chernozem (black earth) areas, the shift of agriculture to the boreal forest zone would not compensate for crop losses due to increasing aridity in South European regions of Russia with the best soils (Dronin and Kirilenko, 2011). Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE 23-4 HERE Figure 23-4: Percentage change in simulated water-limited yield for winter wheat in 2030 with respect to the 2000 baseline for the A1B scenario using ECHAM5 (left column) and HadCM3 (right) GCMs. Upper maps to do not take adaptation into account. Bottom maps include adaptation. Source: Donatelli et al., 2012.] With generally warmer and drier conditions, deep rooted weeds (Gilgen et al., 2010b) and weeds with contrasting physiology, such as C4 species, could pose a more serious threat (Bradley et al., 2010) to crops than shallow rooted C3 weeds (Stratonovitch, 2012). Arthropod-borne diseases (viruses and phytoplasmas), winter infection root and stem diseases (phoma stem canker of oilseed rape and eyespot of wheat) (Butterworth et al., 2010)(West et al., 2012), Fusarium blight (Madgwick et al., 2011), grapevine moth (Caffarra et al., 2012) and a black rot fungus in fruit trees (Weber, 2009) could create increasing damages in Europe under climate change. However, other pathogens like cereal stem rots (e.g. Puccinia striiformis) (Luck et al., 2011) and grapevine powdery mildew (Caffarra et al., 2012) could be limited by increasing temperatures. Increased damages from plant pathogens and insect pests are projected by 2050 in Nordic countries which have hitherto been protected by cold winters and geographic isolation (Hakala et al., 2011; Roos et al., 2011). Some pests, like the European corn borer (Trnka et al., 2007), could also extend their climate niche in Central Europe. Pests and disease management will be affected with regard to timing, preference and efficacy of chemical and biological measures of control (Kersebaum et al., 2008). Autonomous adaptation by farmers, through the advancement of sowing and harvesting dates and the use of longer cycle varieties (Howden et al., 2007; Moriondo et al., 2011; Moriondo et al., 2010; Olesen et al., 2011) could result in a general improvement of European wheat yields in the 2030s compared to the 2000s (Donatelli et al., 2012) (Figure 23-4). However, farmer sowing dates seem to advance slower than crop phenology (Menzel et al., 2006)(Siebert and Ewert, 2012), possibly because earlier sowing is often prevented by lack of soil workability and frost-induced soil crumbling (Oort et al., 2012). Simulation studies which anticipate on earlier sowing in Europe may thus be overly optimistic. Further adaptation options include: changes in crop species, fertilization, irrigation, drainage, land allocation and farming system (Bindi and Olesen, 2011). At the high range of the projected temperature changes, only plant breeding aimed at increasing yield potential jointly with drought resistance and adjusted agronomic practices may reduce risks of yield shortfall (Olesen et al., 2011)(Rötter et al., 2011)(Ventrella et al., 2012). Crop breeding is, however, challenged by temperature and rainfall variability, since: i) breeding has not yet succeeded in altering crop plant development responses to short-term changes in temperature (Parent and Tardieu, 2012) and ii) distinct crop drought tolerance traits are required for mild and severe water deficit scenarios (Tardieu, 2012). Adaptation to increased climatic variability may require an increased use of between and within species genetic diversity in farming systems (Smith and Olesen, 2010) and the development of insurance products against weather-related yield variations (Musshoff et al., 2011). Adaptive capacity and long term economic viability of farming systems may vary given farm structural change induced by climate change (Mandryk et al., 2012); (Moriondo et al., 2010b). In Southern Europe, the regional welfare loss caused by changes in the agriculture sector under a high warming scenario (+5.4°C) was estimated at 1% of GDP. Northern Europe was the single sub-region with welfare gains (+0.7%) from agriculture in this scenario (Ciscar et al., 2011). 23.4.2. Livestock Production Livestock production is adversely affected by heat (Tubiello et al., 2007)(AR5 WG2 7.2.1.3). With intensive systems, heat stress reduced dairy production and growth performance of large finishing pigs at daily mean air temperatures above 18 and 21°C, respectively (André et al., 2011; Renaudeau et al., 2011). High temperature and air humidity during breeding increased cattle mortality risk by 60% in Italy (Crescio et al., 2010). Adaptation requires changes in diets and in farm buildings (Renaudeau et al., 2012) as well as targeted genetic improvement programmes (Hoffmann, 2010). With grass based livestock systems, model simulations (A1B scenario, ensemble of downscaled GCMs) show by end of century increases in potential dairy production in Ireland and France, however with higher risks of summer- autumn production failures in Central Europe and at French sites (Graux et al., 2012; Trnka et al., 2009). Climate conditions projected for the 2070s in central France (A2 scenario) reduced significantly grassland production in a four years experiment under elevated CO2 (Cantarel et al., 2013). At the same site, a single experimental summer Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 drought altered production during the next two years (Zwicke et al., 2013). Resilience of grassland vegetation structure was observed to prolonged experimental heating and water manipulation (Grime et al., 2008). However, weed pressure from tap-rooted forbs was increased after severe experimental summer droughts (Gilgen et al., 2010a). Mediterranean populations could be used to breed more resilient and better adapted forage plant material for livestock production (Poirier et al., 2012). Climate change has affected animal health in Europe [high confidence]. The spread of bluetongue virus in sheep across Europe has been partly attributed to climate change (Arzt et al., 2010)(Guis et al., 2012) through increased seasonal activity of the Culicoides vector (Wilson and Mellor, 2009). The distribution of this vector is unlikely to expand but its abundance could increase in Southern Europe (Acevedo et al., 2010). Ticks, the primary arthropod vectors of zoonotic diseases in Europe (e.g. Lyme disease and tick-borne encephalitis), have changed distributions towards higher altitudes and latitudes with climate change (van Dijk et al., 2010)(Petney et al., 2012; Randolph and Rogers, 2010)(AR5 WG2 23.5). Exposure to fly strike could increase in a warmer climate but adaptation in husbandry practices would limit impacts on livestock (Wall and Ellse, 2011). The overall risk of incursion of Crimean-Congo haemorrhagic fever virus in livestock through infected ticks introduced by migratory bird species would not be increased by climate change (Gale et al., 2012). The probability of introduction and large-scale spread of Rift Valley Fever in Europe is also very low (Chevalier et al., 2010). Epidemiological surveillance and increased coordinated regional monitoring and control programmes have the potential to reduce the incidence of vector-borne animal diseases (Chevalier et al., 2010) (Wilson and Mellor, 2009). 23.4.3. Water Resources and Agriculture Future projected trends confirm the widening of water resource differences between Northern and Southern Europe reported in AR4 (Alcamo et al., 2007). In Southern Europe, soil water content will decline, saturation conditions and drainage will be increasingly rare and restricted to periods in winter and spring, and snow accumulation and melting will change, especially in the mid-mountain areas (García-Ruiz et al., 2011). Across most of Northern and Continental Europe, an increase in flood hazards (Falloon and Betts, 2010)(23.3.1) could increase damages to crops and plant growth, complicate soil workability, and increase yield variability (Olesen et al., 2011). Groundwater recharge and/or water table level would be significantly reduced by the end of the century under A2 scenario for river basins located in Southern Italy, Spain, Northern France and Belgium (Ducharne et al., 2010; Goderniaux et al., 2011; Guardiola-Albert and Jackson, 2011; Senatore et al., 2011). However, non-significant impacts were found for aquifers in Switzerland and in England (Stoll et al., 2011)(Jackson et al., 2011). Less precipitation in summer and higher rainfall during winter could increase nitrate leaching (Kersebaum et al., 2008) with negative impacts on water quality (Bindi and Olesen, 2011). Even with reduced N fertilizer application, groundwater nitrate concentrations would increase by the end of the century in the Seine river basin (Ducharne et al., 2007). More robust water management, pricing and recycling policies, in order to secure adequate future water supply and prevent tensions among users could be required in Southern Europe (García-Ruiz et al., 2011). Reduced suitability for rainfed agricultural production (Daccache and Lamaddalena, 2010; Daccache et al., 2012; Henriques et al., 2008; Trnka et al., 2011) will increase water demand for crop irrigation (Savé et al., 2012). However, increased irrigation may not be a viable option, especially in the Mediterranean area, because of projected declines in total runoff and groundwater resources (Olesen et al., 2011). In a number of catchments water resources are already over-licensed and/or over-abstracted (Daccache et al., 2012) and their reliability is threatened by climate change induced decline in groundwater recharge and to a lesser extent by the increase in potential demand for irrigation (Ducharne et al., 2010; Majone et al., 2012). To match this demand, irrigation system costs could increase by 20-27% in Southern Italy (Daccache and Lamaddalena, 2010) and new irrigation infrastructures would be required in some regions (van der Velde et al., 2010) However, since the economic benefits are expected to be small, the adoption of irrigation would require changes in institutional and market conditions (Finger et al., 2011). Moreover, since aquatic and terrestrial ecosystems are affected by agricultural water use (Klve et al., 2011), irrigation demand restrictions are projected in environmentally focussed future regional scenarios (Henriques et al., 2008). Earlier sowing dates, increased soil organic matter content, low-energy systems, deficit irrigation and improved water use efficiency of irrigation systems and crops can be used as adaptation pathways (Daccache and Lamaddalena, 2010; Gonzalez-Camacho et al., 2008; Lee et al., 2008; Schutze and Schmitz, 2010) especially in Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Southern and south-eastern regions of Europe (Trnka et al., 2009);(Falloon and Betts, 2010). Improved water management in upstream agricultural areas could mitigate adverse impacts downstream (Klve et al., 2011) and groundwater recharge could be targeted in areas with poor water-holding soils (Wessolek and Asseng, 2006). 23.4.4. Forestry Observed and future responses of forests to climate change include changes in growth rates, phenology, composition of animal and plant communities, increased fire and storm damage, and increased insect and pathogen damage. Tree mortality and forest decline due to severe drought events were observed in forest populations in Southern Europe (Affolter et al., 2010; Bigler et al., 2006; Raftoyannis et al., 2008), including Italy (Bertini et al., 2011)(Giuggiola et al., 2010), Cyprus (ECHOES Country report, 2009), and Greece (Raftoyannis et al., 2008) as well as in Belgium (Kint et al., 2012), Switzerland (Rigling et al., 2013) and the pre-Alps in France (Allen et al., 2010; Charru et al., 2010; Rouault et al., 2006). Declines have also been observed in wet forests not normally considered at risk of drought (Choat et al., 2012). An increase in forest productivity has been observed in Russia (Sirotenko and Abashina, 2008). Future projections show that in Northern and Atlantic Europe the increasing atmospheric CO2 and higher temperatures are expected to increase forest growth and wood production, at least in the short-medium term (Lindner et al., 2010). On the other hand, in Southern and eastern Europe, increasing drought and disturbance risks will cause adverse effects and productivity is expected to decline (Hlásny et al., 2011; Keenan et al., 2011; Lavalle et al., 2009; Lindner et al., 2010; Silva et al., 2012; Sirotenko and Abashina, 2008). By 2100, climate change is expected to reduce the economic value of European forest land depending on interest rate and climate scenario, which equates to potential damages of several hundred billion Euros (Hanewinkel et al., 2013). In Southern Europe, fire frequency and wildfire extent significantly increased after the 1970s compared with previous decades (Pausas and Fernández-Munoz, 2012) due to fuel accumulation (Koutsias et al., 2012), climate change (Lavalle et al., 2009) and extreme weather events (Camia and Amatulli, 2009; Carvalho et al., 2011; Hoinka et al., 2009; Koutsias et al., 2012; Salis et al., 2013) especially in the Mediterranean basin (Marques et al., 2011; Pausas and Fernández-Munoz, 2012)(Fernandes et al., 2010; Koutsias et al., 2012). The most severe events in France, Greece, Italy, Portugal, Spain, and Turkey in 2010 were associated with strong winds during a hot dry period (EEA, 2010c). However, for the Mediterranean region as a whole, the total burned area has decreased since 1985 and the number of wildfires has decreased from 2000 to 2009, with large inter-annual variability (Marques et al., 2011; San-Miguel-Ayanz et al., 2012; Turco et al., 2013). Megafires, triggered by extreme climate events, had caused record maxima of burnt areas in some Mediterranean countries during last decades (San-Miguel-Ayanz et al., 2013). Future wildfire risk is projected to increase in Southern Europe (Carvalho et al., 2011; Dury et al., 2011; Lindner et al., 2010; Vilén and Fernandes, 2011), with an increase in the occurrence of high fire danger days (Arca et al., 2012; Lung et al., 2012) and in fire season length (Pellizzaro et al., 2010). The annual burned area is projected to increase by a factor of 3 to 5 in Southern Europe compared to the present under the A2 scenario by 2100 (Dury et al., 2011). In Northern Europe, fires are projected to become less frequent due to increased humidity (Rosan and Hammarlund, 2007). Overall, the projected increase in wildfires is likely to lead to a significant increase in greenhouse gas emissions due to biomass burning (Chiriaco et al., 2013; Pausas et al., 2008; Vilén and Fernandes, 2011), even if often difficult to quantify (Chiriaco et al., 2013). [INSERT FIGURE 23-5 HERE Figure 23-5: Changes in forest fire risk in Europe for two time periods: baseline (left) and 2041 2070 (right), based on high-resolution regional climate models and the SRES A1B emission scenario. Source: Lung et al., 2013.] Wind storm damage to forests in Europe has recently increased (Usbeck et al., 2010). Boreal forests will become more vulnerable to autumn/early spring storm damage due to expected decrease in period of frozen soil (Gardiner et al., 2010). Increased storm losses by 8-19% under A1B and B2 scenarios respectively is projected in Western Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Germany for 2060-2100 compared to 1960-2000, with the highest impacts in the mountainous regions (Klaus et al., 2011; Pinto et al., 2010). An increase in the incidence of diseases has been observed in many European forests (FAO, 2008b; Marcais and Desprez-Loustau, 2007). In Continental Europe, some species of fungi benefit from milder winters and others spread during drought periods from south to north (Drenkhan et al., 2006; Hanso and Drenkhan, 2007). Projected increased late summer warming events will favour diffusion of bark beetle in Scandinavia, in lowland parts of central Europe and Austria (Jönsson et al., 2011; Jönsson et al., 2009)(Seidl et al., 2009). Possible response approaches to the impacts of climate change on forestry include short-term and long-term strategies that focus on enhancing ecosystem resistance and resilience and responding to potential limits to carbon accumulation (Millar et al., 2007; Nabuurs et al., 2013). Fragmented small-scale forest ownership can constrain adaptive capacity (Lindner et al., 2010). Landscape planning and fuel load management may reduce the risk of wildfires but may be constrained by the higher flammability due to warmer and drier conditions (Moreira et al., 2011). Strategies to reduce forest mortality include preference of species better adapted to relatively warm environmental conditions (Resco et al., 2007). The selection of tolerant or resistant families and clones may also reduce the risk of damage by pests and diseases in pure stands (Jactel et al., 2009). 23.4.5. Bioenergy Production The potential distribution of temperate oilseeds (e.g. oilseed rape, sunflower), starch crops (e.g. potatoes), cereals (e.g. barley) and solid biofuel crops (e.g. sorghum, Miscanthus) is projected to increase in Northern Europe by the 2080s, due to increasing temperatures, and to decrease in Southern Europe due to increased drought frequency) (Tuck et al., 2006). Mediterranean oil and solid biofuel crops, currently restricted to Southern Europe, are likely to extend further north (Tuck et al., 2006). The physiological responses of bioenergy crops, in particular C3 Salicaceae trees, to rising atmospheric CO2 concentration may increase drought tolerance due to improved plant water use, consequently yields in temperate environments may remain high in future climate scenarios (Oliver et al., 2009). A future increase in the northward extension of the area for short rotation coppice (SRC) cultivation leading to greenhouse gas neutral is expected (Liberloo et al., 2010). However, the northward expansion of SRC would erode the European terrestrial carbon sink due to intensive management and high turnover of SRC compared to conventional forest where usually harvesting is less than annual growth (Liberloo et al., 2010). 23.4.6. Fisheries and Aquaculture In AR4, Easterling et al. (2007) reported that the recruitment and production of marine fisheries in the North Atlantic are likely to increase. In European seas, warming causes a displacement to the north and/or in depth of fish populations (Daufresne et al., 2009) (AR5 WG2 Chapter 6, 23.6.4) which has a direct impact on fisheries (Cheung et al., 2010; Cheung et al., 2013; Tasker, 2008). For instance, in British waters, the lesser sandeel (Ammodytes marinus), which is a key link in the food web, shows declining recruitments since 2002 and is projected to further decline in the future with a warming climate (Heath et al., 2012). In the Baltic Sea, although some new species would be expected to immigrate because of an expected increase in sea temperature, only a few of these species would be able to successfully colonize the Baltic because of its low salinity (Mackenzie et al., 2007). In response to climate change and intensive fishing, widespread reductions in fish body size (Daufresne et al., 2009) and in the mean size of zooplankton (Beaugrand and Reid, 2012) have been observed over time and these trends further affect the sustainability of fisheries (Pitois and Fox, 2006)(Beaugrand and Kirby, 2010) [see also Chapter 6]. Aquaculture can be affected as the areal extent of some habitats that are suitable for aquaculture can be reduced by sea-level rise. Observed higher water temperatures have adversely affected both wild and farmed freshwater salmon production in the southern part of the distribution areas (Jonsson and Jonsson, 2009). In addition, ocean acidification may disrupt the early developmental stages of shellfish (Callaway et al., 2012). Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Numerous studies confirm the amplification through fishing of the effects of climate change on population dynamics and consequently on fisheries (Planque et al., 2010). The decline of the North Sea cod during the 1980-2000 period results from the combined effects of overfishing and of an ecosystem regime shift due to climate change (Beaugrand and Kirby, 2010). Over the next decade, this stock was not restored from its previous collapse (Mieszkowska et al., 2009)(ICES, 2010). In North Sea and Celtic Seas, the steep decline in boreal species (Henderson, 2007) was compensated for by the arrival of southern (Lusitanian) species (Engelhard et al., 2011; Lenoir et al., 2011; ter Hofstede et al., 2010). Climate change may reinforce parasitic diseases and impose severe risks for aquatic animal health [See Chapter 6]. As water temperatures increase, a number of endemic diseases of both wild and farmed salmonid populations are likely to become more prevalent and threats associated with exotic pathogens may rise (Marcos-Lopez et al., 2010). In Iberian Atlantic, the permitted harvesting period for the mussel aquaculture industry was reduced because of harmful algal blooms resulting from changes in phytoplankton communities linked to a weakening of the Iberian upwelling (Perez et al., 2010). With freshwater systems, summer heat waves boost the development of harmful cyanobacterial blooms (Johnk et al., 2008). For oysters in France, toxic algae may be linked to both climate warming and direct anthropogenic stressors (Buestel et al., 2009). Fishery management thresholds will have to be reassessed as the ecological basis on which existing thresholds have been established changes, and new thresholds will have to be developed for immigrant species (Mackenzie et al., 2007)(Beaugrand and Reid, 2012). These changes may lead to loss of productivity, but also the opening of new fishing opportunities, depending on the interactions between climate impacts, fishing grounds and fleet types. They will also affect fishing regulations, the price of fish products and operating costs, which in turn will affect the economic performance of the fleets (Cheung et al., 2012). Climate change impacts on fisheries profits range from negative for sardine fishery in the Iberian Atlantic fishing grounds (Perez et al., 2010)(Garza-Gil et al., 2010), to non-significant for the Bay of Biscay (Le Floc'h et al., 2008) and positive on the Portuguese coast, since most of the immigrant fish species are marketable (Vinagre et al., 2011). Human social fishing systems dealing with high variability upwelling systems with rapidly reproducing fish species may have greater capacities to adjust to the additional stress of climate change than human social fishing systems focused on longer-lived and generally less variable species (Perry et al., 2011; Perry et al., 2010). Climate change adaptation is being considered for integration in European maritime and fisheries operational programs (European Commission, 2013). _____ START BOX 23-1 HERE _____ Box 23-1. Assessment of Climate Change Impacts on Ecosystem Services by Sub-Region Ecosystems provide a number of vital provisioning, regulating and cultural services for people and society that flow from the stock of natural capital (Stoate et al., 2009)(Harrison et al., 2010). Provisioning services such as food from agro-ecosystems or timber from forests derive from intensively managed ecosystems; regulating services underpin the functioning of the climate and hydrological systems; and, cultural services such as tourism, recreation and aesthetic value are vital for societal well-being (see section 23.5.4). The table summarises the potential impacts of climate change on ecosystem services in Europe by sub-region based on an assessment of the published literature (2004-2013). The direction of change (increasing, decreasing or neutral) is provided, as well as the number of studies/papers on which the assessment was based (in brackets). Empty cells indicate the absence of appropriate literature. Unless otherwise stated, impacts assume no adaptation and are assessed for the mid-century (2050s). A decrease in natural hazard regulation (e.g. for wildfires) implies an increased risk of the hazard occurring. Biodiversity is included here as a service (for completeness), although it is debated whether biodiversity should be considered as a service or as part of the natural capital from which services flow. What is agreed, however, is that biodiversity losses within an ecosystem will have deleterious effects on service provision (Mouillot et al., 2013). The provision of ecosystem services in Southern Europe is projected to decline across all service categories in response to climate change [high confidence]. Other European sub-regions are projected to have both losses and gains in the provision of ecosystem services [high confidence]. The Northern sub-region will have increases in provisioning services arising from climate change [high confidence]. Except for the Southern sub-region, the effects of climate change on regulating services are balanced with respect to gains and losses [high confidence]. There are Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 fewer studies for cultural services, although these indicate a balance in service provision for the Alpine and Atlantic regions, with decreases in service provision for the Continental, Northern and Southern sub-regions [low confidence]. [INSERT BOX 23-1 TABLE HERE] _____ END BOX 23-1 HERE _____ 23.5. Implications of Climate Change for Health and Social Welfare 23.5.1. Human Population Health Climate change is likely to have a range of health effects in Europe. Further studies since AR4 have confirmed the effects of heat on mortality and morbidity in European populations and particularly in older people and those with chronic disease (Corobov et al., 2012; Corobov et al., 2013; Kovats and Hajat, 2008; Aström et al., 2011). With respect to sub-regional vulnerability, populations in southern Europe appear to be most sensitive to hot weather (Baccini et al., 2011; D'Ippoliti et al., 2010; Michelozzi et al., 2009; Michelozzi et al., 2009), and also will experience the highest heat exposures (Figure 23-2). However, populations in Continental (Hertel et al., 2009) and Northern Europe (Rocklöv and Forsberg, 2010)(Armstrong et al., 2011)(Varakina et al., 2011) are also vulnerable to heat wave events. Adaptation measures to reduce heat health effects include heat wave plans (Bittner et al., 2013) which have been shown to reduce heat-related mortality in Italy (Schifano et al., 2012), but evidence of effectiveness is still very limited (Hajat et al., 2010; Lowe et al., 2011). There is little information about how future changes in housing and infrastructure (23.3.2) would reduce the regional or local future burden of heat-related mortality or morbidity. Climate change is likely to increase future heat-related mortality (Baccini et al., 2011; Ballester et al., 2011; Huang et al., 2011) and morbidity (Aström et al., 2013), although most published risk assessments do not include consideration of adaptation (Huang et al., 2011). For most countries in Europe, the current burden of cold-related mortality (Analitis et al., 2008) is greater than the burden of heat mortality. Climate change is likely to reduce future cold-related mortality (Ballester et al., 2011; HPA, 2012)(AR5 WG2 11.4.1). Mortality and morbidity associated with flooding is becoming better understood although the surveillance of health effects of disasters remains inadequate (WHO, 2013). Additional flood mortality due to sea level rise has been estimated in the Netherlands (Maaskant et al., 2009); and in the UK for river flooding (Hames and Vardoulakis, 2012) but estimates of future mortality due to flooding are highly uncertain. There remains limited evidence regarding the long term mental health impacts of flood events (Paranjothy et al., 2011; WHO, 2013). Evidence about future risks from climate change with respect to infectious diseases is still limited (Randolph and Rogers, 2010; Semenza and Menne, 2009; Semenza et al., 2012). There have been developments in mapping the current and potential future distribution of important disease vector species in Europe. The Asian tiger mosquito Aedes albopictus (a vector of dengue and Chikungunya (Queyriaux et al., 2008) is currently present in Southern Europe (ECDC, 2009) and may extend eastward and northward under climate change (Caminade et al., 2012; Fisher et al., 2011; Roiz et al., 2011). The risk of introduction of dengue remains very low because it would depend upon the introduction and expansion of the Ae. Aegypti together with the absence of effective vector control measures (ECDC, 2012). Climate change is unlikely to affect the distribution of visceral and cutaneous leishmaniasis (currently present in the Mediterranean region) in the near term (Ready, 2010). However, in the long term (15-20 years), there is potential for climate change to facilitate the expansion of either vectors or current parasites northwards (Ready, 2010). The risk of introduction of exotic Leishmania species was considered very low due to the low competence of current vectors (Fischer et al., 2010a). The effect of climate change on the risk of imported or locally-transmitted (autochthonous) malaria in Europe has been assessed in Spain (Sainz-Elipe et al., 2010), France (Linard et al., 2009) and the UK (Lindsay et al., 2010). Disease re-emergence would depend upon many factors including: the introduction of a large population of infectious people or mosquitoes, high levels of people-vector contact, resulting from significant changes in land use, as well as climate change (see chapter 11). Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Since AR4, there is more evidence on implications of climate change on food safety at all stages from production to consumption (FAO, 2008a; Jacxsens et al., 2010; Popov Janevska et al., 2010). The sensitive of salmonellosis to temperature has declined in recent years (Lake et al., 2009) and the overall incidence of salmonellosis is declining in most European countries (Semenza et al., 2012). Climate change may also have affects on food consumption patterns. Weather affects pre and post harvest mycotoxin production but the implications of climate change are unclear. Cold regions may become liable to temperate-zone problems concerning contamination ochratoxin A, patulin and Fusarium toxins (Paterson and Lima, 2010). A control of the environment of storage facilities may avoid post-harvest problems but at additional cost (Paterson and Lima, 2010). Other potential consequences concern marine biotoxins in seafood following production of phycotoxins by harmful algal blooms and the presence of pathogenic bacteria in foods following more frequent extreme weather conditions (Miraglia et al., 2009). There is little evidence that climate change will affect human exposures to contaminants in the soil or water (e.g. persistent organic pollutants). Risk modelling is often developed for single exposure agents (e.g. a pesticide) with known routes of exposure. These are difficult to scale up to the population level. The multiple mechanisms by climate may affect transmission or contamination routes also makes this very complex (Boxall et al., 2009). Adaptation in the health sector has so far been largely limited to the development of heat health warning systems, but many research gaps regarding effective adaptation options (HPA, 2012). A survey of national infectious disease experts in Europe identified several institutional changes that needed to be addressed to improve future responses to climate change risks: ongoing surveillance programs, collaboration with veterinary sector and management of animal disease outbreaks, national monitoring and control of climate-sensitive infectious diseases, health services during an infectious disease outbreak and diagnostic support during an epidemic (Semenza et al., 2012). 23.5.2. Critical Infrastructure Critical national infrastructure is defined as the assets (physical or electronic) that are vital to the continued delivery and integrity of the essential services upon which a country relies, the loss or compromise of which would lead to severe economic or social consequences or to loss of life. Extreme weather events, such as floods, heat waves and wild fires are known to damage critical infrastructure. The UK floods in 2007 led to significant damage to power and water utilities, and to communications and transport infrastructure (Chatterton et al., 2010) (Table 23-1). Forest fires can affect transport infrastructure, as well as the destruction of buildings. Major storms in Sweden and Finland have led to loss of trees, with damage to the power distribution network, leading to electricity blackouts lasting weeks, as well as the paralysis of services such as rail transport and other public services that depend on grid electricity. Health system infrastructure (hospitals, clinics) is vulnerable to extreme events, particularly flooding (Radovic et al., 2012). The heat waves of 2003 and 2006 had adverse effects on patients and staff in hospitals from overheating of buildings. Evidence from France and Italy indicate that death rates in in-patients increased significantly during heat wave events (Ferron et al., 2006; Stafoggia et al., 2008). Further, higher temperatures have had serious implications for the delivery of healthcare, as well drug storage and transport (Carmichael et al., 2013). 23.5.3. Social Impacts There is little evidence regarding the implications of climate change for employment and/or livelihoods in Europe. However, the evidence so far (as reviewed in this chapter) indicates that there are likely to be changes to some industries (e.g. tourism, agriculture) that may lead to changes in employment opportunities by region and by sector. Current damages from weather-related disasters (floods and storms) are significant (23.3.1). Disasters have long lasting effects of the affected populations (Schnitzler et al., 2007). Households are often displaced while their homes are repaired (Whittle et al., 2010). Little research has been carried out on the impact of extreme weather events such Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 as heat waves and flooding on temporary or permanent displacement in Europe. Coastal erosion associated with sea level rise, storm surges and coastal flooding will require coastal retreat in some of Europe s low lying areas (Philippart et al., 2011). Managed retreat is also an adaptation option in some coastal areas. Concerns have been raised about equality of access to adaptation within coastal populations at risk from climate change. For example, a study in the UK found that vulnerability to climate change in coastal communities is likely to be increased by social deprivation (Zsamboky et al., 2011). In the European region, the indigenous populations are present in Arctic regions are considered vulnerable to climate change impacts on livelihoods and food sources (Arctic Climate Impact Assessment, 2005) [12.3.4, 28.2.4]. Research has focussed on indigenous knowledge, impacts on traditional food sources and community responses/adaptation (Mustonen and Mustonen, 2011a; Mustonen and Mustonen, 2011b). However, these communities are also experiencing rapid social, economic and other non-climate-related environmental changes (such as oil and gas exploration) [see 28.2.4]. There is evidence the climate change has altered the seasonal behaviour of pastoralist populations, such as the Nenets reindeer herders in northern Russia (Amstislavski et al., 2013). However, socio-economic factors may be more important than climate change for the future sustainability of Reindeer husbandry (Rees et al., 2008) [28.2.3.5]. 23.5.4. Cultural Heritage and Landscapes Climate change will affect culturally-valued buildings (Storm et al., 2008) through extreme events and chronic damage to materials (Brimblecombe et al., 2006; Brimblecombe and Grossi, 2010; Brimblecombe, 2010a; Brimblecombe, 2010b; Grossi et al., 2011)(Sabbioni et al., 2012). Cultural heritage is a non-renewable resource and impacts from environmental changes are assessed over long timescales (Brimblecombe and Grossi, 2008)(Bonazza et al., 2009a; Bonazza et al., 2009b; Brimblecombe and Grossi, 2009; Brimblecombe and Grossi, 2010; Grossi et al., 2008). Climate change may also affect indoor environments where cultural heritage is preserved (Lankester and Brimblecombe, 2010) as well as visitor behaviour at heritage sites (Grossi et al., 2010). There is also evidence to suggest that climate change and sea level rise will affect maritime heritage in the form of shipwrecks and other submerged archaeology (Björdal, 2012). Surface recession on marble and compact limestone will be affected by climate change (Bonazza et al., 2009a). Marble monuments in Southern Europe will continue to experience high levels of thermal stress (Bonazza et al., 2009b) but warming is likely to reduce frost damage across Europe, except in Northern and Alpine Europe and permafrost areas (Iceland) (Grossi et al., 2007; Sabbioni et al., 2008). Damage to porous materials due to salt crystallisation may increase all over Europe (Benavente et al., 2008; Grossi et al., 2011). In Northern and Eastern Europe, wood structures will need additional protection against rainwater and high winds (Sabbioni et al., 2012). AR4 concluded that current flood defences would not protect Venice from climate change. Venice now has a flood forecasting system, and is introducing the MOSE system of flood barriers (Keskitalo, 2010). Recent evidence suggests, however, that climate change may lead to a decrease in the frequency of extreme storm surges in this area (Troccoli et al., 2012a). Europe has many unique rural landscapes, which reflect the cultural heritage that has evolved from centuries of human intervention, e.g. the cork oak based Montado in Portugal, the Garrigue of southern France, Alpine meadows, grouse moors in the UK, machair in Scotland, peatlands in Ireland, the polders of Belgium and the Netherlands and vineyards. Many, if not all, of these cultural landscapes are sensitive to climate change and even small changes in the climate could have significant impacts (Gifford et al., 2011). Alpine meadows, for example, are culturally important within Europe, but although there is analysis of the economics (tourism, farming) and functionality (water run-off, flooding and carbon sequestration) of these landscapes there is very little understanding of how climate change will affect the cultural aspects on which local communities depend. Because of their societal value, cultural landscapes are often protected and managed through rural development and environmental policies. The peat-rich uplands of northern Europe, for example, have begun to consider landscape management as a means of adapting to the effects of climate change (e.g. the moors for the future partnership in the Peak District National Park, UK). For a discussion of the cultural implications of climate change for vineyards see Box 23-2. Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ START BOX 23-2 HERE _____ Box 23-2. Implications of Climate Change for European Wine and Vineyards Wine production in Europe accounts for more than 60% of the global total (Goode, 2012) and makes an important contribution to cultural identity. Apart from impacts on grapevine yield, higher temperatures are also expected to affect wine quality in some regions and grape varieties by changing the ratio between sugar and acids (Bock et al., 2011)(Santos et al., 2011)(Duchene et al., 2010). In western and central Europe, projected future changes could benefit wine quality, but might also demarcate new potential areas for viticulture (Malheiro et al., 2010). Adaptation measures are already occurring in some vineyards (e.g. vine management, technological measures, production control and to a smaller extent relocation) (Battaglini et al., 2009; Duarte Alonso and O Neill, 2011; Holland and Smit, 2010; Malheiro et al., 2010; Moriondo et al., 2011; Santos et al., 2011).Vineyards may be displaced geographically beyond their traditional boundaries ( terroir linked to soil, climate and traditions) (Metzger and Rounsevell, 2011), and in principle, wine producers could adapt to this problem by growing grape varieties that are more suited to warmer climates. Such technical solutions, however, do not account for the unique characteristics of wine production cultures and consumer perceptions of wine quality that strongly affect the prices paid for the best wines (Metzger and Rounsevell, 2011)(White et al., 2009). It would become very difficult, for example, to produce fine wines from the cool-climate Pinot Noir grape within its traditional terroir of Burgundy under many future climate scenarios, but consumers may not be willing to pay current day prices for red wines produced from other grape varieties (Metzger and Rounsevell, 2011). An additional barrier to adaptation is that wine is usually produced within rigid, regionally-specific, regulatory frameworks that often prescribe, amongst other things, what grapes can be grown where, e.g., the French AOC or the Italian DOC and DOCG designations. Suggestions have been made to replace these rigid concepts of regional identity with a geographically flexible terroir that ties a historical or constructed sense of culture to the wine maker and not to the region (White et al., 2009). _____ END BOX 23-2 HERE _____ 23.6. Implications of Climate Change for the Protection of Environmental Quality and Biological Conservation Terrestrial and freshwater ecosystems provide a number of vital services for people and society, such as biodiversity, food, fibre, water resources, carbon sequestration and recreation (Box 23-1). 23.6.1. Air Quality Climate change will have complex and local effects on pollution chemistry, transport, emissions and deposition. Outdoor air pollutants have adverse effects on human health, biodiversity, crop yields and cultural heritage. The main outcomes of concern are both the average (background) levels and peak events for tropospheric ozone, particulates, sulphur oxides (SOx) and nitrogen oxides (NOx). Future pollutant concentrations in Europe have been assessed using atmospheric chemistry models, principally for ozone (Forkel and Knoche, 2006; Forkel and Knoche, 2007). Reviews have concluded that GCM/CTM studies find that climate change per se (assuming no change in future emissions or other factors) is likely to increase summer tropospheric ozone levels (range 1 10 ppb) by 2050s in polluted areas (that is, where concentrations of precursor nitrogen oxides are higher) (AQEG, 2007; Jacob and Winner, 2009)[see also 21.4.1.3.2.]. The effect of future climate change alone on future concentrations of particulates, nitrogen oxides and volatile organic compounds is much more uncertain. Higher temperatures also affect natural emissions volatile organic compounds (VOCs) which are ozone precursors (Hartikainen et al., 2012). One study has projected an increase in fire-related air pollution (O3 and PM10) in Southern Europe (Carvalho et al., 2011). Overall, the model studies are inconsistent regarding future projections of background level and exceedences. Recent evidence has shown adverse impacts on agriculture from even low concentrations of ozone, however, there is more consistent evidence now regarding the threshold for health (mortality) impacts of ozone. Therefore, it is Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 unclear whether increases in background levels below health-related thresholds would be associated with an increased burden of ill health. Some studies have attributed an observed increase in European ozone levels to observed warming (Meleux et al., 2007), which appears to be driven by the increase in extreme heat events in 2003, 2006 and 2010 (Solberg et al., 2008). Peak ozone events were observed during the major heat waves in Europe in multiple countries. Wildfire events have had an impact on local and regional on air quality (Hodzic et al., 2007; Liu et al., 2009; Miranda et al., 2009) which have implications for human health (Analitis et al., 2012) (Table 23-1). 23.6.2. Soil Quality and Land Degradation The current cost of soil erosion, organic matter decline, salinisation, landslides and contamination is estimated to be EUR 38 billion annually for the EU (JRC-EEA, 2010), in the form of damage to infrastructures, treatment of water contaminated through the soil, disposal of sediments, depreciation of land and costs related to the ecosystem functions of soil (JRC-EEA, 2010). Projections show significant reductions in summer soil moisture in the Mediterranean region, and increases in the north-eastern part of Europe (Calanca et al., 2006). Climate change impacts on erosion shows diverging evidence under the A2 scenario. In Tuscany, even with a decline in precipitation volume until 2070, in some month higher erosion rates would occur due to higher rainfall erosivity (Marker et al., 2008). For two Danish river catchments, assuming a steady-state land use, suspended sediment transport would increase by 17-27% by 2071-2100 (Thodsen et al., 2008; Thodsen, 2007). In Upper-Austria, with the regional climate model HadRM3H, a small reduction in average soil losses is projected for croplands in all tillage systems, however with high uncertainty (Scholz et al., 2008). In Northern Ireland, erosion decreases are generally projected with downscaled GCMs for a case study hillslope (Mullan et al., 2012). Adaptive land-use management can reduce the impact of climate change through soil conservation methods like zero tillage and conversion of arable to grasslands (Klik and Eitzinger, 2010). In central Europe, compared to conventional tillage, conservation tillage systems reduced modelled soil erosion rates under future climate scenarios by between 49 and 87% (Scholz et al., 2008). Preserving upland vegetation reduced both erosion and loss of soil carbon and favoured the delivery of a high quality water resource (House et al., 2011); (McHugh, 2007). Maintaining soil water retention capacity, e.g. through adaptation measures (Post et al., 2008), contributes to reduce risks of flooding as soil organic matter absorbs up to twenty times its weight in water. 23.6.3. Water Quality Climate change may affect water quality in several ways, with implications for food production and forestry (23.4.3), ecosystem functioning (Box 23-1), human and animal health, and compliance with environmental quality standards, including those of the Water Framework Directive. Shallower waters will witness a more rapid temperature increase than deeper waters, since heat is absorbed mainly in the upper water layers and turbulent mixing is truncated by shallow depth. In parallel, a decrease in saturating oxygen concentrations occurs. Since AR4, there is further evidence of adverse effects caused by extreme weather events: reductions in dissolved oxygen, algal blooms (Mooij et al., 2007; Ulén and Weyhenmeyer, 2007) during hot weather, and contamination of surface and coastal waters with sewage and/or chemicals (pesticides) after rainfall (Boxall et al., 2009). A reduction in rainfall may lead to low flows which increase concentrations of biological and chemical contaminants. Reduced drainage can also enhance sedimentation in drainage systems and hence enhance particle-bound phosphorous retention and reduce phosphorous load to downstream higher order streams (Hellmann and Vermaat, 2012). Variability in changes in rainfall and run-off, as well as water temperature increases, will lead to differences in water quality impacts by sub-region. Climate change is projected to increase nutrient loadings: in Northern Europe this is caused by increased surface runoff; in Southern Europe this is caused by increased evapotranspiration and increased concentrations due to reduced volumes of receiving lakes (Jeppesen et al., 2011). Local studies generally confirm this pattern: increased nutrient loads are foreseen in Danish watersheds (Andersen et al., 2006), France (Delpla et al., 2011) and the UK (Howden et al., 2010; Macleod et al., 2012; Whitehead et al., 2009); AR5 WG2 Chapter Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 4.3.2.5). In larger rivers, such as the Meuse, increased summer temperature and drought can lead to more favourable conditions for algal blooms and reduced dilution capacity of effluent from industry and sewage works (van Vliet and Zwolsman, 2008). 23.6.4. Terrestrial and Freshwater Ecosystems Current and future climate changes including CO2 increase are determining negative effects of habitat loss on species density and diversity (Mantyka-pringle et al., 2012)(Rickebusch et al., 2008). Projected habitat loss is greater for species at higher elevations (Engler et al., 2011)(Castellari, 2009; Dullinger et al., 2012) and suitable habitats for Europe's breeding birds are projected to shift nearly 550 km northeast by the end of the century (Huntley et al., 2007). Aquatic habitats and habitat connectivity in river networks may become increasingly fragmented (Blaustein et al., 2010; Della Bella et al., 2008; Elzinga et al., 2007; Gómez-Rodríguez et al., 2010; Hartel et al., 2011; Morán-López et al., 2012)(Harrison et al., 2008)(Clark et al., 2010a)(Clark et al., 2010; Fronzek et al., 2006; Fronzek et al., 2010; Fronzek et al., 2011; Gallego-Sala et al., 2010). Despite some local successes and increasing responses, the rate of biodiversity loss does not appear to be slowing (Butchart et al., 2010). The effectiveness of Natura 2000 areas to respond to climate change has been questioned (Araújo et al., 2011). However, when considering connectivity related to the spatial properties of the network, the Natura 2000 network appears rather robust (Mazaris et al., 2013). Several studies now highlight the importance of taking into account climate change projections in the selection of conservation areas (Araújo et al., 2011; Ellwanger et al., 2011; Filz et al., 2013; Virkkala et al., 2013). Observed changes in plant communities in European mountainous regions show a shift of species ranges to higher altitudes resulting in species richness increase in boreal-temperate mountain regions and decrease in Mediterranean mountain regions (Pauli et al., 2012)(Gottfried et al., 2012). In Southern Europe, a great reduction in phylogenetic diversity of plant, bird and mammal assemblages will occur, and gains are expected in regions of high latitude or altitude for 2020, 2050 and 2080. However, losses will not be offset by gains and a trend towards homogenization across the continent will be observed (Thuiller et al., 2011)(Alkemade et al., 2011). Large range contractions due to climate change are projected for several populations of Pinus cembra and Pinus Sylvestris (Casalegno et al.,2010)(Giuggiola et al., 2010) while for the dominant Mediterranean tree species, Holm oak, a substantial range expansion is projected under A1B emissions scenario (Cheaib et al., 2012). The human impacts on distribution of tree species landscape may make them more vulnerable to climate change (del Barrio et al., 2006; Hemery et al., 2010). Observed climate changes are altering breeding seasons, timing of spring migration, breeding habitats, latitudinal distribution and migratory behaviour of birds (Feehan et al., 2009) (Jonzén et al., 2006; Rubolini et al., 2007a; Rubolini et al., 2007b)(Lemoine et al., 2007a; Lemoine et al., 2007b). A northward shift in bird community composition has been observed (Devictor et al., 2008). Common species of European birds with the lowest thermal maxima have showed the sharpest declines between 1980 and 2005 (Jiguet et al., 2010). Projections for 120 native terrestrial non-volant European mammals suggest that 5-9% are at risk of extinction, assuming no migration, during the 21st century due to climate change, while 70-78% may be severely threatened under A1 and B2 climatic scenarios (Levinsky et al., 2007). Those populations not showing a phenological response to climate change may decline (Moller et al., 2008), such as amphibian and reptile species (Araújo et al., 2006), or experience ecological mismatches (Saino et al., 2011). Climate change can affect trophic interactions, as co- occurring species may not react in a similar manner. Novel emergent ecosystems composed of new species assemblages arising from differential rates of range shifts of species can occur (Keith et al., 2009; Montoya and Raffaelli, 2010; Schweiger et al., 2012). Since invasive alien species rarely change their original climatic niches (Petitpierre et al., 2012), climate change can exacerbate the threat posed by invasive species to biodiversity in Europe (West et al., 2012) amplifying the effects of introduction of the exotic material such as alien bioenergy crops (EEA, 2012), pest and diseases (Aragon and Lobo, 2012), tropical planktonic species (Cellamare et al., 2010) and tropical vascular plants (Skeffington and Hall, 2011; Taylor et al., 2012). Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 23.6.5. Coastal and Marine Ecosystems Climate change will affect Europe s coastal and marine ecosystems by altering the biodiversity, functional dynamics and ecosystem services of coastal wetlands, dunes, inter-tidal and subtidal habitats, offshore shelves, seamounts and currents (Halpern et al., 2008) through changes in eutrophication, invasive species, species range shifts, changes in fish stocks and habitat loss (Doney et al., 2011)(EEA, 2010d). The relative magnitude of these changes will vary temporally and spatially, requiring a range of adaptation strategies that target different policy measures, audiences and instruments (Philippart et al., 2011)(Airoldi and Bec, 2007). Europe s northern seas are experiencing greater increases in sea surface temperatures (SSTs) than the southern seas, with the Baltic, North and Black seas warming at 2-4 times the mean global rate (Philippart et al., 2011)(Belkin, 2009). In the Baltic, decreased sea ice will expose coastal areas to more storms, changing the coastal geomorphology (BACC, 2008)(HELCOM, 2007). Warming SSTs will influence biodiversity and drive changes in depth and latitudinal range for intertidal and sub-tidal marine communities, particularly in the North and Celtic seas (Hawkins et al., 2011)(Sorte et al., 2010)(Wethey et al., 2011). Warming is affecting food chains and changing phenological rates (Durant et al., 2007). For example, changes in the timing and location of phytoplankton and zooplankton are affecting North Sea cod larvae (Beaugrand et al., 2010)(Beaugrand and Kirby, 2010). Temperature changes have affected the distribution of fisheries in all seas over the past 30 years (Beaugrand and Kirby, 2010)(Hermant et al., 2010). Warmer waters also increase the rate of the establishment and spread of invasive species, further altering trophic dynamics and the productivity of coastal marine ecosystems (Molnar et al., 2008)(Rahel and Olden, 2008). Changes in the semi-enclosed seas could be indicative of future conditions in other coastal-marine ecosystems (Lejeusne et al., 2009). In the Mediterranean, invasive species have arrived in recent years at the rate of one introduction every 4 to 5 weeks (Streftaris et al., 2005). While in this case the distribution of endemic species remained stable, most non-native species have spread northward by an average of 300 km since the 1980s, resulting in an area of spatial overlap with invasive species replacing natives by nearly 25% in 20 years. Dune systems will be lost in some places due to coastal erosion from combined storm surge and sea level rise, requiring restoration (Day et al., 2008)(Ciscar et al., 2011)(Magnan et al., 2009). In the North Sea, the Iberian coast, and Bay of Biscay, a combination of coastal erosion, infrastructure development and sea defences may lead to narrower coastal zones ( coastal squeeze ) (EEA, 2010d)(Jackson and McIlvenny, 2011)(OSPAR, 2010). 23.7. Cross-Sectoral Adaptation Decision-making and Risk Management Studies on impacts and adaptation in Europe generally consider single sectors or outcomes, as described in the previous sections of this chapter. For adaptation decision-making, more comprehensive approaches are required. Considerable progress has been made to advance planning and development of adaptation measures, including the economic analyses (Section 23.7.6) (see Box 23-3), and the developed of climate services (Medri et al., 2012; WMO, 2011). At the international level, the European Union has started adaptation planning, through information sharing (Climate-ADAPT platform) and legislation (EC, 2013a). National and local governments are also beginning to monitor progress on adaptation, including the development of a range of indicators (UK-ASC, 2011). _____ START BOX 23-3 HERE _____ Box 23-3. National and Local Adaptation Strategies The increasing number of national (EEA, 2013) and local (Heidrich et al., 2013) adaptation strategies in Europe has led to research on their evaluation and implementation (Biesbroek et al., 2010b). Many adaptation strategies were found to be agendas for further research, awareness raising and/or coordination and communication for implementation (e.g. (Dumollard and Leseur, 2011; Pfenniger et al., 2010). Actual implementation often was limited Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 to disaster risk reduction, environmental protection, spatial planning (23.7.4), and coastal zone and water resources management. The implementation of planned adaptation at the national level was attributed to political will and good financial and information capacity (Westerhoff et al., 2011). Analysis of seven national adaptation strategies (Denmark, Finland, France, Germany, Netherlands, Spain, UK) found that that while there is a high political commitment to adaptation planning and implementation, evaluation of the strategies and actual implementation is yet to be defined (Biesbroek et al., 2010a; Swart et al., 2009b; Westerhoff et al., 2011). One of the earliest national adaptation strategies (Finland) has been evaluated, in order to compare identified adaptation measures with those launched in different sectors. It has found that while good progress has been made on research and identification of options, few measures have been implemented except in the water resources sector (Ministry of Agriculture and Forestry, 2009). At the local government level, adaptation plans are being developed in several cities (EEA, 2013), including London (GLA, 2010), Madrid, Manchester, Copenhagen, Helsinki, and Rotterdam. Adaptation in general is a low priority for many European cities, and many plans do not have adaptation priority as the main focus (Carter, 2011). Many studies are covering sectors sensitive to climate variability, as well as sectors that are currently under pressure from socioeconomic development. A recent assessment found a lack of cross-sector impact and adaptation linkages as an important weakness in the city plans (Hunt and Watkiss, 2011). Flexibility in adaptation decision making needs to be maintained (Hallegatte et al., 2008)(Biesbroek et al., 2010b). _____ END BOX 23-3 HERE _____ 23.7.1. Coastal Zone Management Coastal zone management and coastal protection plans that integrate adaptation concerns are now being implemented. Underlying scientific studies increasingly assess effectiveness and costs of specific options (Hilpert et al., 2007)(Kabat et al., 2009)(Dawson et al., 2011) (23.7.6). Early response measures are needed for floods and coastal erosion, to ensure that climate change considerations are incorporated into marine strategies, with mechanisms for regular updating (OSPAR, 2010; UNEP, 2010). In the Dutch plan for flood protection (Delta Committee, 2008), adaptation to increasing river runoff and sea level rise plays a prominent role. It also includes synergies with nature conservation and fresh water storage (Kabat et al., 2009), and links to urban renovation (cost estimates are included in Section 23.7.6). While that plan mostly relies on large scale measures, new approaches such as small-scale containment of flood risks through compartmentalisation are also studied (Klijn et al., 2009). The UK government has developed extensive adaptation plans (TE2100) to adjust and improve flood defences for the protection of London from future storm surges and flooding (Environmental Agency, 2009). An elaborate analysis has provided insight in the pathways for different adaptation options and decision-pointss that will depend on the eventual sea-level rise. 23.7.2. Integrated Water Resource Management Water resources management in Europe has experienced a general shift from hard to soft measures that allow more flexible responses to environmental change (Pahl-Wostl, 2007). Integrated water resource management explicitly includes the consideration of environmental and social impacts (Wiering and Arts, 2006). Climate change has been incorporated into water resources planning in England and Wales (Arnell, 2011)(Charlton and Arnell, 2011)(Wade et al., 2013) and in the Netherlands (de Graaff et al., 2009). The robustness of adaptation strategies for water management in Europe has been tested in England (Dessai and Hulme, 2007) and Denmark (Haasnoot et al., 2012; Refsgaard et al., 2013). Other studies have emphasised the search for robust pathways, for instance in the Netherlands (Haasnoot et al., 2012; Kwadijk et al., 2010). Public participation has also increased in decision making, e.g. river basin management planning (Huntjens et al., 2010), flood defence plans (e.g. TE2100), and drought contingency plans (Iglesias et al., 2007). Guidance has been developed on the inclusion of adaptation in water management (UNECE, 2009) and river basin management plans (EC, 2009b). Adaptation in the water sector could also be achieved through the EU Water Framework and Flood Directives (Quevauviller, 2011), but a study of Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 decision makers, including local basin managers, identified several important barriers to this (Brouwer et al., 2013). Water allocation between upstream and downstream countries is challenging in regions exposed to prolonged droughts such as the Euphrates-Tigris river basin, where Turkey plans to more than double water abstraction by 2023(EEA, 2010a). 23.7.3. Disaster Risk Reduction and Risk Management A series of approaches to disaster risk management are employed in Europe, in response to national and European policy developments to assess and reduce natural hazard risks. New developments since the AR4 include assessment and protection efforts in accordance with the EU Floods Directive (European Parliament and Council, 2007), the mapping of flood risks, improve civil protection response and early warning systems (Ciavola et al., 2011). Most national policies address hazard assessment and do not include analyses of possible impacts (de Moel et al., 2009). The effectiveness has been assessed of flood protection (Bouwer et al., 2010) and also non-structural or household level measures to reduce losses from river flooding (Botzen et al., 2010a)(Dawson et al., 2011). Some studies show that current plans may be insufficient to cope with increasing risks from climate change, as shown for instance for the Rhine river basin (Te Linde et al., 2010a; Te Linde et al., 2010b). Other options that are being explored are the reduction of consequences, response measures, and increasing social capital (Kuhlicke et al., 2011), as well as options for insuring and transferring losses (Section 23.3.7). The Netherlands carried out a large-scale analysis and simulation exercise to study the possible emergency and evacuation response for a worst-case flood event (ten Brinke et al., 2010). Increasing attention is also being paid in Europe to non-government actions that can reduce possible impacts from extreme events. Terpstra and Gutteling (2008) found through a survey that individual citizens are willing to assume some responsibility for managing flood risk, and they are willing to contribute to preparations in order to reduce impacts. Survey evidence is available for Germany and the Netherlands that, under certain conditions, individuals can be encouraged to adopt loss prevention measures (Thieken et al., 2006)(Botzen et al., 2009). Small businesses can reduce risks when informed about possibilities immediately after an event (Wedawatta and Ingirige, 2012). 23.7.4. Land Use Planning Spatial planning policies can build resilience to the impacts of climate change (Bulkeley, 2010). However, the integration of adaptation into spatial planning is often limited to a general level of policy formulation that can sometimes lack concrete instruments and measures for implementation in practice (Mickwitz et al., 2009)(Swart et al., 2009a). There is evidence to suggest the widespread failure of planning policy to account for future climate change (Branquart et al., 2008). Furthermore, a lack of institutional frameworks to support adaptation is, potentially, a major barrier to the governance of adaptation through spatial planning (ESPACE, 2007)[chapter 16]. Climate change adaptation is often treated as a water management or flooding issue, which omits other important aspects of the contribution of land use planning to adaptation (Mickwitz et al., 2009)(Wilson, 2006)(Van Nieuwaal et al., 2009). For example, in the UK, houses were still being built in flood risk areas (2001-2011) due to competing needs to increase the housing stock (ARUP, 2011). City governance is also dominated by the issues of climate mitigation and energy consumption rather than in adapting to climate change (Bulkeley, 2010; Heidrich et al., 2013). Some cities, e.g. Rotterdam, have started to create climate adaptation plans and this process tends to be driven by the strong political leadership of mayors (Sanchez-Rodriguez, 2009). The Helsinki Metropolitan Area s Climate Change Adaptation Strategy (HSY, 2010) is a regional approach focusing on the built environment in the cities of Helsinki, Espoo, Vantaa and Kauniainen, and their surroundings. It includes approaches for dealing with increasing heat waves, more droughts, milder winters, increasing (winter) precipitation, heavy rainfall events, river floods, storm surges, drainage water floods and sea level rise. Green infrastructure provides both climate adaptation and mitigation benefits as well as offering a range of other benefits to urban areas, including health improvements, amenity value, inward investment, and the reduction of Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 noise and outdoor air pollution. Green infrastructure is an attractive climate adaptation option since it also contributes to the sustainable development of urban areas (Gill et al., 2007; James et al., 2009). Urban green space and green roofs can moderate temperature and decrease surface rainwater run-off (Gill et al., 2007). Despite the benefits however of urban green space, conflict can occur between the use of land for green space and building developments (Hamin and Gurran, 2009). European policies for biodiversity (e.g. the European Biodiversity Strategy (EC, 2011)) look to spatial planning to help protect and safeguard internationally and nationally designated sites, networks and species, as well as locally valued sites in urban and non-urban areas, and to create new opportunities for biodiversity through the development process (Wilson, 2008). Conservation planning in response to climate change impacts on species aims to involve several strategies to better manage isolated habitats, increase colonisation capacity of new climate zones and optimise conservation networks to establish climate refugia (Vos et al., 2008). 23.7.5. Rural Development Rural development is one of the key policy areas for Europe, yet there is little or no discussion about the role of climate change in affecting future rural development. The EU White Paper on adapting to climate change (EC, 2009a) encourages Member States to embed climate change adaptation into the three strands of rural development aimed at improving competitiveness, the environment, and the quality of life in rural areas. It appears however that little progress has been made in achieving these objectives. For example, the EUs Leader programme was designed to help rural actors improve the long-term potential of their local areas by encouraging the implementation of sustainable development strategies. Many Leader projects address climate change adaptation, but only as a secondary or in many cases a non-intentional by-product of the primary rural development goals. The World Bank s community adaptation project has seen a preponderance of proposals from rural areas in Eastern Europe and Central Asia (Heltberg et al., 2012) suggesting that adaptation based development needs in Eastern Europe are currently not being met by policy. 23.7.6. Economic Assessments of Adaptation Compared to studies assessed in AR4 (AR4 WG2 Chapter 17.2.3), costs estimates for Europe are increasingly derived from bottom-up and sector-specific studies, aimed at costing response measures (Watkiss and Hunt, 2010), in addition to the economy-wide assessments (Aaheim et al., 2012). The evidence base, however, is still fragmented and incomplete. The coverage of adaptation costs and benefit estimates is dominated by structural (physical) protection measures, where effectiveness and cost components can be more easily identified. For energy, agriculture, infrastructure there is medium coverage of cost and benefit categories. There is a lack of information regarding adaptation costs in the health and social care sector. Table 23-2 summarises some of the more comprehensive cost estimates for Europe for sectors at regional and national level. It is stressed that the costing studies use a range of methods and metrics and relate to different time periods and sectors, which renders robust comparison difficult. As an example, there are large differences between the cost estimates for coastal and river protection in the Netherlands and other parts of Europe (Table 23-2), this is due to the objectives for adaptation and the large differences in the level of acceptable risk. For example, Rojas et al. (2013) assess a 1 in 100 year level of protection for Europe, while the Netherlands has set standards up to 1 in 4,000 and 10,000 year level return periods. More detailed treatment of the economics of adaptation is provided in AR5 WG2 Chapter 17. [INSERT TABLE 23-2 HERE Table 23-2: Selected published cost estimates for planned adaptation in European countries.] Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 23.7.7. Barriers and Limits to Adaptation The implementation of adaptation options presents a range of opportunities, constraints and limits. Constrains (barriers) to implementation are financial, technical and political (see discussion in AR5 WG2 16). Some impacts will be unavoidable due to limits (physical, technological, social, economic or political). Examples of limits in the European context are described by sector in Table 23-3. For example, the contraints on building or extending flood defences would include pressure for land, conservation needs, and amenity value of coastal areas (AR5 WG2 5.5.5). Towards the end of the century, it is likely that adaptation limits are expected to be reached earlier under higher rates of warming. Opportunities and co-benefits of adaptation are also discussed in section 23.8 below. [INSERT TABLE 23-3 HERE Table 23-3: Limits to adaptation to climate change.] 23.8. Co-Benefits and Unintended Consequences of Adaptation and Mitigation Scientific evidence for decision making is more useful if impacts are considered in the context of impacts on other sectors and in relation to adaptation, mitigation and other important policy goals. The benefits of adaptation and mitigation policies can be felt in the near term and in the local population, although benefits relating to greenhouse gas emissions reduction may not be apparent until the longer term. The benefits of adaptation measures are often assessed using conventional economic analyses, some of which include non-markets costs and benefits (externalities)(Watkiss and Hunt, 2010). This section will describe policies, strategies and measures where there is good evidence regarding mitigation/adaptation costs and benefits. Few studies have quantified directly the trade- offs/synergies for a given policy. 23.8.1. Production and Infrastructure Mitigation policies (decarbonisation strategies) are likely to have important implications for dwellings across Europe. The unintended consequences of mitigation in the housing sector include: changes to household energy prices and adverse effects from decreased ventilation in dwellings (Mavrogianni et al., 2012)(Davies and Oreszczyn, 2012)(Jenkins et al., 2008; Jenkins, 2009). The location, type and dominant energy use of the building will determine its overall energy gain or loss to maintain comfort levels. Adaptation measures such as the use of cooling devices will probably increase a building s energy consumption if no other mitigation measures are applied. The potential for cooling dwellings without increased energy consumption, and with health benefits is large (Wilkinson et al., 2009). When looking at the broader context of urban infrastructures, despite existing efforts to include both adaptation, and mitigation into sustainable development strategies at city level (e.g. Hague, Rotterdam, Hamburg, Madrid, London, Manchester), priority on adaptation still remains low (Carter, 2011). There is potential to develop strategies that can address both mitigation and adaptation solutions, as well as have health and environmental benefits (Milner et al., 2012). In energy supply, the adverse effect of climate change on water resources in some coastal regions in southern Europe may further enhance the development of desalination plants as an adaptation measure, possibly increasing energy consumption and thus greenhouse gases emissions. Coastal flood defence measures may alter vector habits and have implications for local vector-borne disease transmission (Medlock and Vaux, 2013). In tourism, adaptation and mitigation may be antagonistic, as in the case of artificial snowmaking in European ski resorts which requires significant amounts of energy and water (OECD, 2007; Rixen et al., 2011) and the case of desalination for potable water production which also requires energy. However, depending on the location and size of the resort, implications are expected to differ and thus need to be investigated on a case-by-case basis. A similar relationship between adaptation and mitigation may hold for tourist settlements in southern Europe, where expected temperature increases during the summer may require increased cooling in order to maintain tourist comfort and thus increase greenhouse gas emissions and operating costs. Furthermore, a change of tourist flows as a result of Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 tourists adapting to climate change may affect transport emissions, while mitigation in transport could also lead to a change in transport prices and thus possibly affect tourist flows. 23.8.2. Agriculture, Forestry, and Bioenergy Agriculture and forestry face two challenges under climate change, both to reduce emissions and to adapt to a changing and more variable climate (Smith and Olesen, 2010)(Lavalle et al., 2009). The agriculture sector contributes to about 10% of the total anthropogenic greenhouse gas (GHG) emissions in the EU27 (EEA, 2010b). Estimates of European carbon dioxide, methane and nitrous oxide fluxes between 2000 and 2005 suggest that methane emissions from livestock and nitrous oxide emissions from agriculture are fully compensated for by the carbon dioxide sink provided by forests and by grassland soils (Schulze et al., 2010). However, projections following a baseline scenario suggest a significant decline (-25 to -40%) of the forest carbon sink of the EU until 2030 compared to 2010. Using wood for bioenergy results initially in a carbon debt due to reduced storage in forests, which affects the net GHG balance depending on the energy type that is replaced and the time span considered (McKechnie et al., 2011). Including additional bioenergy targets of EU member states has an effect on the development of the European forest carbon sink (and on the carbon stock), which is not accounted for in the EU emission reduction target (Bottcher et al., 2012). In arable production systems, adapting to climate change by increasing the resilience of crop yields to heat and to rainfall variability would have positive impacts on mitigation by reducing soil erosion, as well as soil organic carbon and nitrogen losses. Improving soil water holding capacity through the addition of crop residues and manure to arable soils, or by adding diversity to the crop rotations, may contribute both to adaptation and to mitigation (Smith and Olesen, 2010). There are also synergies and trade-offs between mitigation and adaptation options for soil tillage, irrigation and livestock breeding (Smith and Olesen, 2010). Reduced tillage (and no-till) may contribute to both adaptation and mitigation as it tends to reduce soil erosion and run-off (Soane et al., 2012) and fossil-fuel use (Khaledian et al., 2010), while increasing in some situations soil organic carbon stock (Powlson et al., 2011). However, increased N2O emission may negate the mitigation effect of reduced tillage (Powlson et al., 2011). Irrigation may enhance soil carbon sequestration in arable systems (Rosenzweig et al., 2008)(Rosenzweig and Tubiello, 2007), but increased irrigation under climate change would increase energy use and may reduce water availability for hydro-power (reduced mitigation potential) (Wreford et al., 2010). In intensive livestock systems, warmer conditions in the coming decades might trigger the implementation of enhanced cooling and ventilation in farm buildings (Rosenzweig and Tubiello, 2007), thereby increasing energy use and associated GHG emissions. In grass-based livestock systems, adaptation by adjusting the mean annual animal stocking density to the herbage growth potential (Graux et al., 2012) is likely to create a positive feedback on GHG emissions per unit area (Soussana and Luscher, 2007; Soussana et al., 2010). Land management options may also create synergies and trade-offs between mitigation and adaptation. Careful adaptation of forestry and soil management practices will be required to preserve a continental ecosystem carbon sink in Europe (Schulze et al., 2010) despite the vulnerability of this sink to climatic extremes (Ciais et al., 2005) and first signs of carbon sink saturation in European forest biomass (Nabuurs et al., 2013). In areas that are vulnerable to extreme events (e.g. fires, storms, droughts) or with high water demand, the development of bioenergy production from energy crops and from agricultural residues (De Wit et al., 2011) (Fischer et al., 2010b) could further increase demands on adaptation (Wreford et al., 2010). Conversely, increased demands on mitigation could be induced by the potential expansion of agriculture at high latitudes which may release large amounts of carbon and nitrogen from organic soils (Rosenzweig and Tubiello, 2007). 23.8.3. Social and Health Impacts Significant research has been undertaken since AR4 on the health co-benefits of mitigation policies (see WGIII chapters on Housing, Transport and Energy, and WGII chapter 11). Several assessment have quantified benefits in terms of lives saved by reducing particulate air pollution, and trying to coherent policy objectives for emissions Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 reductions in local and global pollution. Policies that improve health from changes in transport and energy can be said to have a general benefit to population health and resilience (Haines et al., 2009a; Haines et al., 2009b). Changes to housing and energy policies also have indirect implications for human health. Researches on the benefits of various housing options (including retrofitting) have been intensively addressed in the context of low energy, healthy and sustainable housing (see WGIII). 23.8.4. Environmental Quality and Biological Conservation There are several conservation management approaches that can address both mitigation, adaptation and biodiversity objectives (Lal et al., 2011). Some infrastructure adaptation strategies, such as desalinisation, sea defences and flood control infrastructure may have negative effects on both mitigation and biodiversity. However, approaches, such as forest conservation and urban green space (23.7.4) have multiple benefits and potentially significant effects. There has been relatively little research about the impacts of future land use demand for bioenergy production, food production and urbanisation on nature conservation. Figure 23-6 (Paterson et al., 2008) summarizes the evidence regarding mitigation and adaptation options on biodiversity assessed from the literature. The figure shows that the options that come closest to being win-win-win are green rooftops, urban tree planting, forest conservation and low-till cultivation. Other options with clear benefits are afforestation, forest pest control, increased farmland irrigation and species translocation. [INSERT FIGURE 23-6 HERE Figure 23-6: Adaptation and mitigation options and their effects on biodiversity. The horizontal axis ranges from positive effects on biodiversity (left-hand side) to negative effects (right-hand side). Each mitigation/adaptation option is located on the biodiversity effect axis (solid bars), including an estimate of the uncertainties associated with the assessment (error bars). The various options are given vertically with mitigation at the top and adaptation at the bottom. Options located toward the centre of the vertical axis have benefits for both mitigation and adaptation. Thus, options located at the centre left of the figure have benefits for mitigation, adaptation and biodiversity and hence are labelled as win-win-win . Other combinations of benefits and dis-benefits are labelled accordingly, e.g. win-lose-win, lose-win-lose, etc. Based on Paterson et al., 2009.] 23.9. Synthesis of Key Findings 23.9.1. Key Vulnerabilities Climate change will have adverse impacts in nearly all sectors and across all sub-regions. Table 23-4 describes the range of impacts projected in 2050 on infrastructure, settlements, environmental quality and the health and welfare of the European population. The projected impacts of climate change on ecosystem services (including food production) are described in Box 23-1. A key finding is that all sub-regions are vulnerable to some impacts from climate change but that these impacts differ significantly in type between the sub-regions. Impacts in neighbouring regions (inter-regional) may also redistribute economic activities across the European landscape. The sectors most likely to be affected by climate change, and therefore with implications for economic activity and population movement (changes in employment opportunities) include: tourism (23.3.6), agriculture (23.4.1), and forestry (23.4.4). [INSERT TABLE 23-4 HERE Table 23-4: Assessment of climate change impacts by sub-region by 2050, assuming a medium emissions scenario, and no planned adaptation. Impacts assume economic development, including land use change. Impacts are assessed for the whole sub-region, although differences in impact within sub-regions are estimated for some impacts.] The majority of published assessments are based on climate projections in the range 1-4 degrees global mean temperature per century. Under these scenarios, regions in Europe may experience higher rates of warming (in the Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 range 4-6 degrees per century), due to climate variability (Jacob et al. 2013). Limited evidence exists on the potential impacts in Europe under very high rates of warming (>4 degrees above pre-industrial levels) but these would lead to large increase in coastal flood risk as well as impacts on global cereal yields and other effects on the global economy (AR5 WG2 19.5.1). Many key vulnerabilities are already well known since the AR4, but some new vulnerabilities are emerging based on the evidence reviewed in this report. The policy/governance context in Europe is extremely important in determining key vulnerabilities (either reducing or exacerbating vulnerability) since Europe is a highly regulated region. Further, vulnerability will be strongly affected by changes in the non-climate drivers of change (e.g. economic, social protection measures, governance, technological drivers). Extreme events affect multiple sectors and have the potential to cause a systemic impacts from secondary effects (chapter 19). Past events indicate the vulnerability of transport, energy, agriculture, water resources and health systems. Resilience to very extreme events varies by sector, and by country (Ludwig et al., 2011; Pitt, 2008; Ulbrich et al., 2012). Extreme events (heat waves and droughts) have had significant impacts on populations as well impacts on multiple economic sectors (Table 23-1), and resilience to future heat waves has only been addressed within some sectors. However, there is surprisingly little evidence regarding the impacts of major extreme events (e.g. Russian heat wave of 2010) and on responses implemented post-event to increase resilience. Future vulnerability will also be strongly affected by cross-sectoral (indirect) interactions, e.g. flooding-ecosystems, agriculture-species, agriculture- cultural landscapes, and so on. Climate change is likely to have significant impacts on future water availability, and the increased risks of water restrictions in Southern, Central and Atlantic sub-regions. Studies indicate a significant reduction in water availability from river abstraction and from groundwater resources, combined to increased demands from a range of sectors (irrigation, energy and industry, domestic use) and to reduced water drainage and run-off (as a result of increased evaporative demand) (Ludwig et al., 2011). Climate change will affect rural landscapes by modifying relative land values, and hence competition, between different land-uses (Smith et al., 2010). This will occur directly, e.g. through changes in the productivity of crops and trees [23.4], and indirectly through climate change impacts on the global supply of land-based commodities and their movement through international trade [23.9.2]. Climate change will have a range of impacts in different European sub-regions. The adaptive capacity of populations is likely to vary significantly within Europe. Adaptive capacity indicators have been developed based on future changes in socio-economic indicators and projections (Lung et al., 2012; Metzger et al., 2008)(Acosta et al., 2013). These studies concluded that the Nordic countries have higher adaptive capacity than most of the Southern European countries, with countries around the Mediterranean having a lower capacity than the countries around the Baltic Sea region. Some regions or areas are particularly vulnerable to climate change: Populations and infrastructure in coastal regions are likely to be adversely affected by sea level rise, particularly after mid-century [23.3.1, 23.5.3]. Urban areas are also vulnerable due to high density of people and built infrastructure from weather extremes [23.3, 23.5.1]. High mountains. Due to high impact of climate change on natural hazard, water and snow resources and lack of migration possibilities for plant species, mountain regions concentrate vulnerabilities in infrastructure for transport and energy sectors, as well as for tourism, agriculture and biodiversity Mediterranean region will suffer multiple stresses and systemic failures due to climate changes. Changes in species composition, increase of alien species, habitat losses and degradation both in land and sea together with agricultural and forests production losses due to increasing heat waves and droughts exacerbated also by the competition for water will increase the sub-region vulnerability (Ulbrich et al., 2012). The following risks have emerged from observations of climate sensitivity and observed adaptation: Arable crop yields. There is new evidence to suggest that crop yields and production may be more vulnerable as a result of increasing climate variability. This will limit the potential poleward expansion of agricultural production. Limits to genetic progress to adapt are increasingly reported. Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 New evidence regarding implications during summer on inland waterways (decreased access) and long range ocean transport (increased access). Terrestrial and freshwater species are vulnerable from climate-change shifts in habitats. There is new evidence that species cannot populate new habitat due to habitat fragmentation (urbanization). Observed migration rates are less than that assumed in modelling studies. There are legal barriers to introducing new species (e.g. forest species in France). New evidence that phenological mismatch will cause additional adverse effects on some species. A positive (and emerging) effect that may reduce vulnerability is that many European governments (and individual cities) have become aware of the need to adapt to climate change and so are developing and/or implementing adaptation strategies and measures. Additional risks have emerged from the assessed literature: Increased summer energy demand, especially in southern Europe, requires additional power generation capacity, which will be under-utilised during the rest of the year, entailing higher supply costs. Housing will be affected, with increased overheating under no adaptation and damage from subsidence and flooding. Passive cooling measures alone are unlikely to be sufficient to address adaptation in all regions and types of buildings. Retrofitting current housing stock will be expensive. An emerging concern is the vulnerability of cultural heritage, including monuments/buildings and cultural landscapes. Some cultural landscapes will disappear. Grape production is highly sensitive to climate, but production (of grape varieties) is strongly culturally-dependent and adaptation is potentially limited by the regulatory context. Good evidence that climate change will increase distribution and seasonal activity of pests and diseases. Limited evidence that such effects already occurring. Increased threats to plant and animal health. Public policies are in place to reduce pesticide use in agriculture use and antibiotics in livestock, and this will increase vulnerability to the impact of climate change on agriculture and livestock production. Lack of institutional frameworks is a major barrier to adaptation governance. In particularly, the systematic failure in land use planning policy to account for climate change. [INSERT TABLE 23-5 HERE Table 23-5: Key risks from climate change in Europe and the potential for reducing risk through mitigation and adaptation. Risk levels are presented in three timeframes: the present, near-term (2030-2040), and longer-term (2080-2100). For each timeframe, risk levels are estimated for a continuation of current adaptation and for a hypothetical highly adapted state. For a given key risk, change in risk level through time and across magnitudes of climate change is illustrated, but because the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks, sectors, or regions. Key risks were identified based on assessment of the literature and expert judgment.] 23.9.2. Climate Change Impacts Outside Europe and Inter-Regional Implications With increasing globalization, the impacts of climate change outside the European region are likely to have implications for countries within the region. For example, the Mediterranean region (Southern Europe and non- European Mediterranean countries) has been considered high vulnerable to climate change (Navarra, 2013). Eastern European countries have, in general, lower adaptive capacity than Western or Northern European countries. The high volume of international travel increases Europe s vulnerability to invasive species, including the vectors of human and animal infectious diseases. The transport of animals and animal products has facilitated the spread of animal diseases (Conraths and Mettenleiter, 2011). Important exotic vectors that have become established in Europe include the vector Aedes albopictus (Becker, 2009) (23.5.1). Another inter-regional implication concerns the changes in the location of commercial fish stocks shared between countries. Such changes may render existing international agreements regarding the sharing of yield from these stocks obsolete giving rise to international disputes (Arnason, 2012). For instance, the North Sea mackerel stock has recently been extending westwards beyond the EU jurisdiction into the Exclusive Economic Zones of Iceland and Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 the Faroe Islands, which unilaterally claimed quota for mackerel Territorial disagreements of this type could increase in the future with climate change. Although several studies have proposed a role of climate change to increase migration pressures in low and middle income countries in the future, there is little robust information regarding the role of climate change, environmental resource depletion and weather disasters in future inter-continental population movements. The effect of climate change on external migration flows into Europe is highly uncertain (see chapter 12.4.1 for a more complete discussion). Modelling future migration patterns is complex and so far no robust approaches have been developed. 23.9.3. Effects of Observed Climate Change in Europe Table 23-6 summarises the evidence with respect to key indicators in Europe for the detection of a trend and the attribution of that trend to observed changes in climate factors. The attribution of local warming to anthropogenic climate change is less certain (see Chapter 18 for a full discussion). Further and better quality evidence since 2007 supports the conclusion of AR4 (Europe chapter, Alcamo et al., 2007) that climate change is affecting land, freshwater and marine ecosystems in Europe. Observed warming has caused advancement in the life cycles of many animal groups, including frogs spawning, birds nesting and the arrival of migrant birds and butterflies (see WGII chapter 4 and review by (Feehan et al., 2009). There is further evidence that observed climate change is already affecting agricultural, forest and fisheries productivity (see 23.4). The frequency of river flood events, and annual flood and windstorm damages in Europe have increased over recent decades, but this increase is mainly due to increased exposure and the contribution of observed climate change is unclear (high confidence based on robust evidence and high agreement)(SREX 4.5.3, (Barredo, 2010). The observed increase in the frequency of hot days and hot nights (high confidence, WGI) is likely to have increased heat-related health effects in Europe (medium confidence), and well as a decrease in cold related health effects (medium confidence) (Christidis et al., 2010). Multiple impacts on health, welfare and economic sectors were observed due to the major heat wave events of 2003 and 2010 in Europe (Table 23-5) (see Chapter 18 for discussion on attribution of events). [INSERT TABLE 23-6 HERE Table 23-6: Observed changes in key indicators in ecological and human systems attributable to climate factors.] 23.9.4. Key Knowledge Gaps and Research Needs There is a clear mismatch between the volume of scientific work on climate change since the AR4 and the insights and understanding required for policy needs, as many categories of impacts are still understudied. Some specific research needs have been identified: Little information is available on integrated and cross-sectoral climate change impacts in Europe, as the impact studies typically describe a single sector [see sections 23.3 to 23.6]. This also includes a lack of information on cross-sector vulnerabilities, and the indirect effects of climate change impacts and adaptation responses. This is a major barrier in developing successful evidence-based adaptation strategies that are cost-effective. Climate change impact models are difficult to validate [sections 23.3 to 23.6]; proper testing of the characteristics of baseline impact estimates against baseline information and data would improve their reliability, or the development of alternative methods where baseline data are not available. There is little knowledge on the co-benefits and unintended consequences of adaptation options across a range of sectors [section 23.3 to 23.6]. There is a need to better monitor and evaluate local and national adaptation and mitigation responses to climate change, in both public and private sectors [23.7; Box 23-3]. This includes policies and strategies as well as the effectiveness of individual adaptation measures. Evaluation of adaptation strategies, over a range of time-scales, would better support decision-making. While some means for reporting of national Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 actions exist in Europe (e.g. EU Climate-ADAPT), there is no consistent method of monitoring or a mechanism for information exchange [23.7]. There are now more economic methods and tools available for the costing and valuation of specific adaptation options, in particular for flood defences, water, energy, and agriculture sectors [23.7.6]. However, for other sectors, such as biodiversity, business and industry, and population health costs, cost estimates are still lacking or incomplete. The usefulness of this costing information in decision making need to be evaluated and research can be undertaken to make economic evaluation more relevant to decision making. The need for local climate information to inform decision making also needs to be evaluated. Further research is needed on the effects of climate change on critical infrastructure (including transport, and water and energy supplies, health services) [23.5.2]. Further research is needed on the role of governance in adaptation (including local and national institutions) with respect to implementation of measures in the urban environment, including flood defences, over- heating, and urban planning. The impacts for Europe from high end scenarios of climate change (above 4°C global average warming, with higher temperature change in Europe) are yet unknown. This is because such scenarios have only recently become available, and related impact studies still need to be undertaken for Europe. More study of the implications of climate change for rural development would inform policy in this area [23.7.5]. There is also a lack of information on the resilience of cultural landscapes and communities, and how to manage adaptation, particularly in low technology (productively marginal) landscapes. More research is needed for the medium and long- term monitoring of forest responses and adaptation to climate change and on the predictive modelling of wildfire distribution to better address adaptation and planning policies. There is also a lack of information on the impact of climate changes and climate extremes on carbon sequestration potential of agricultural and forestry systems [23.4.4]. More research on the impacts of climate change on transport, especially on the vulnerability of road and rail infrastructure, and on the contribution of climatic and non-climatic parameters in the vulnerability of air transport (e.g. changes in air traffic volumes, airport capacities, air traffic demand, weather at the airports of origin, intermediate and final destination) [23.3.3]. Improved monitoring of droughts is needed to support the management of crop production [23.4]. Remote sensing could be complemented by field experiments that assess the combined effects of elevated CO2 and extreme heat and drought on crops and pastures. Research is needed on the resilience of human populations to extreme events (factors which increase resilience), including responses to flood and heat wave risks. Inequalities and how adaptation policies may increase or reduce social inequalities [23.5]. Development of improved risk models for vector borne disease (human and animal diseases) in order to support health planning and surveillance [23.4.2, 23.5.1]. A major barrier to research is lack of access to data, which is variable across regions and countries (especially with respect to socio-economic data, climate data, forestry, and routine health data). There is a need for long term monitoring of environmental and social indicators and to ensure open access to data for long-term and sustainable research programmes. Cross-regional cooperation could also ensure compatibility and consistency of parameters across the European region. Frequently Asked Questions FAQ 23.1: Will I still be able to live on the coast in Europe? [to remain at the end of the chapter] Coastal areas affected by storm surges will face increased risk both because of the increasing frequency and of storms and because of higher sea level. Most of this increase in risk will occur after the middle of this century. Models of the coast line suggest that populations in the north western region of Europe are most affected and many countries, including the Netherlands, Germany, France, Belgium, Denmark, Spain and Italy, will need to strengthen their coastal defences. Some countries have already raised their coastal defence standards. The combination of raised sea defences and coastal erosion may lead to narrower coastal zones in the North Sea, the Iberian coast, and Bay of Biscay. Adapting dwellings and commercial buildings to occasional flooding is another response to climate change. Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 But while adapting buildings in coastal communities and upgrading coastal defences can significantly reduce adverse impacts of sea level rise and storm surges, they cannot eliminate these risks, especially as sea levels will continue to rise over time. In some locations, managed retreat is likely to become a necessary response. FAQ 23.2: Will climate change introduce new infectious diseases into Europe? [to remain at the end of the chapter] Many factors play a role in the introduction of infectious diseases into new areas. Factors that determine whether a disease changes distribution include: importation from international travel of people, vectors or hosts (insects, agricultural products), changes in vector or host susceptibility, drug resistance, and environmental changes, such as land use change or climate change. One area of concern that has gained attention is the potential for climate change to facilitate the spread tropical diseases, such as malaria, into Europe. Malaria was once endemic in Europe. Even though its mosquito vectors are still present and international travel introduces fresh cases, malaria has not become established in Europe because infected people are quickly detected and treated. Maintaining good health surveillance and good health systems are therefore essential to prevent diseases from spreading. When an outbreak has occurred (i.e. the introduction of a new disease) determining the causes is often difficult. It is likely that a combination of factors will be important. A suitable climate is a necessary but not a sufficient factor for the introduction of new infectious diseases. FAQ 23.3: Will Europe need to import more food because of climate change? [to remain at the end of the chapter] Europe is one of the world s largest and most productive suppliers of food, but also imports large amounts of some agricultural commodities. A reduction in crop yields, particularly wheat in southern Europe, is expected under future climate scenarios. A shift in cultivation areas of high value crops, such as grapes for wine, may also occur. Loss of food production may be compensated by increases in other European sub-regions. However, if the capacity of the European food production system to sustain climate shock events is exceeded, the region would require exceptional food importation. References Aaheim, A., H. Amundsen, T. Dokken, and T. Wei, 2012: Impacts and adaptation to climate change in European economies. Global Environmental Change, 22(4), 959-968. Aakre, S. and D.T.G. Rübbelke, 2010: Adaptation to climate change in the European Union: efficiency versus equity considerations. Environmental Policy and Governance, 20(3), 159-179. Aakre, S., I. Banaszak, R. Mechler, D. Rübbelke, A. Wreford, and H. Kalirai, 2010: Financial adaptation to disaster risk in the European Union; Identifying roles for the public sector. Mitigation and Adaptation Strategies for Global Change, 15(7), 721-736. ABI, 2009: The Financial Risk of Climate Change. Research Paper No. 19. [Dailey, P., Huddleston, M., Brown, S. and Fasking, D. (eds.)]. Association of British Insurers, London, United Kingdom, pp. 1-107. Acevedo, P., F. Ruiz-Fons, R. Estrada, A.L. Márquez, M.A. Miranda, C. Gortázar, and J. Lucientes, 2010: A broad assessment of factors in determining Culicoides imicola abundance: modelling the present and forecasting its future in climate change scenarios. PloS One, 5(12), e14236. Acosta, L., R.J.T. Klein, P. Reitsma, M.J. Metzger, M.D.A. Rounsevell, R. Leemans, and D. Schroter, 2013: A spatially explicit scenario-driven model of adaptive capacity to global change in Europe. Global Environmental Change, doi.org/10.1016/j.gloenvcha.2013.03.008. Aerts, J., T. Sprong, B. Bannink, J. Bessembinder, E. Koomen, C. Jacobs, N. van der Hoeven, D. Huitema, S. van t Klooster, J. Veraart, A. Walraven, S.N. Jonkman, B. Maaskant, L.M. Bouwer, K. de Bruijn, E. Oosterveld, H. Schuurman, K. Peters, W. Ottevanger, W. Immerzeel, P. Droogers, J. Kwadijk, J. Kind, L. Voogt, H. van der Klis, R. Dellink, F. Affolter, P. Bubeck, M. van der Meulen, G. de Lange, B. Bregman, H. van den Brink, H. Buiteveld, S. Drijfhout, A. Feijt, W. Hazeleger, B. van den Hurk, C. Katsman, A. Kattenberg, G. Lenderink, E. Meijgaard, P. Siegmund, M. de Wit, M. Naples, E. van Velzen, and J. van Zetten, 2008: Aandacht voor Veiligheid. [Aerts, J., Sprong, T. and Bannink, B. (eds.)]. Leven met Water, Klimaat voor Ruimte, DG Water, Netherlands, pp. 1-198. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Affolter, P., U. Büntgen, J. Esper, A. Rigling, P. Weber, J. Luterbacher, and D. Frank, 2010: Inner Alpine conifer response to 20th century drought swings. European Journal of Forest Research, 129, 289-298. AGRESTE, 2011: Agreste Infos rapides-Grandes cultures et fourrages - n°6/7 Les prairies vues par ISOP en septembre 2011. [Cassagne, J.P. (ed.)]. Ministere de l'agriculture, de l'alimentation, de la peche, de la ruralité et de l'aménagement du territoire, Montreuil, France, pp. 1-4. Airoldi, L. and M.W. Bec, 2007: Loss, status and trends for coastal marine habitats of Europe. Oceanography and Marine Biology Annual Review, 45, 345-405. Albertson, K., J. Aylen, G. Cavan, and J. McMorrow, 2010: Climate change and the future occurrence of moorland wildfires in the Peak District of the UK. Climate Research, 45, 105-118. Albrecht, F., T. Wahl, J. Jensen, and R. Weisse, 2011: Determining sea level change in the German Bight. Ocean Dynamics, 61(12), 2037-2050. Alcamo, J., J.M. Moreno, B. Novaky, M. Bindi, R. Corobov, R.J.N. Devoy, C. Giannakopoulos, E. Martin, J.E. Olesen, and A. Shvidenko, 2007: Europe. Climate Change 2007: Impacts, Adaptation and Vulnerability. In: Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Parry, M.L., Canziani, O.F., Palutikof, J.P., van der Linden, P.J. and Hanson, C.E. (eds.)]. Cambridge University Press, Cambridge, United Kingdom, pp. 541-580. Alkemade, R., M. Bakkenes, and B. Eickhout, 2011: Towards a general relationship between climate change and biodiversity: an example for plant species in Europe. Regional Environmental Change, 11, S143-S150. Allen, C.D., A.K. Macalady, H. Chenchouni, D. Bachelet, N. McDowell, M. Vennetier, T. Kitzberger, A. Rigling, D.D. Breshears, E.H.(. Hogg, P. Gonzalez, R. Fensham, Z. Zhang, J. Castro, N. Demidova, J. Lim, G. Allard, S.W. Running, A. Semerci, and N. Cobb, 2010: A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. Forest Ecology and Management, 259(4), 660-684. Amelung, B., S. Nicholls, and D. Viner, 2007: Implications of global climate change for tourism flows and seasonality. Journal of Travel Research, 45, 285-296. Amelung, B. and A. Moreno, 2009: Impacts of climate change in tourism in Europe. PESETA-Tourism study. JRC Scientific and Technical Reports, Seville, Spain, pp. 55. Amelung, B. and A. Moreno, 2012: Costing the impact of climate change on tourism in Europe: results of the PESETA project. Climatic Change, 112(1), 83-100. Amstislavski, P., L. Zubov, H. Chen, P. Ceccato, J.F. Pekel, and I.J. Weedon, 2013: Effects of increase in temperature and open water on transmigration and access to health care by the Nenets reindeer herders in northern Russia. International Journal of Circumpolar Health, 72(suppl 1), Article num. 21183. Amthor, J.S., 2001: Effects of atmospheric CO2 concentration on wheat yield: review of results from experiments using various approaches to control CO2 concentration. Field Crops Research, 73(1), 1-34. Analitis, A., I. Georgiadis, and K. Katsouyanni, 2012: Forest fires are associated with elevated mortality in a dense urban setting. Occupational and Environmental Medicine, 69(3), 158-162. Analitis, A., K. Katsouyanni, A. Biggeri, M. Baccini, B. Forsberg, L. Bisanti, U. Kirchmayer, F. Ballester, E. Cadum, P.G. Goodman, A. Hojs, J. Sunyer, P. Tiittanen, and P. Michelozzi, 2008: Effects of cold weather on mortality: results from 15 European cities within the PHEWE project. American Journal of Epidemiology, 168(12), 1397-1408. Andersen, H.E., B. Kronvang, S. Larsen, C.C. Hoffmann, T.S. Jensen, and E.K. Rasmussen, 2006: Climate-change impacts on hydrology and nutrients in a Danish lowland river basin. Science of the Total Environment, 365(1- 3), 223-237. Andersson, A.K. and L. Chapman, 2011a: The impact of climate change in winter road maintenance and traffic accidents in West Midlands, UK. Accident Analysis and Prevention, 43(1), 284-289. Andersson, A.K. and L. Chapman, 2011b: The use of a temporal analogue to predict future traffic accidents and winter road conditions in Sweden. Meteorological Applications, 18, 125-136. André, G., B. Engel, P.B.M. Berentsen, T. Vellinga, and A.G.J.M. Oude Lansink, 2011: Quantifying the effect of heat stress on daily milk yield and monitoring dynamic changes using an adaptive dynamic model. Journal of Dairy Science, 94(9), 4502-4513. AQEG, 2007: Air Quality and Climate Change: A UK Perspective. Third report of the Air Quality Expert Group. Department for Environment, Food and Rural Affairs - DEFRA, London, United Kingdom, pp. 272. Aragon, P. and J.M. Lobo, 2012: Predicted effect of climate change on the invasibility and distribution of the western corn root-worm. Agricultural and Forest Entomology, 14, 13-18. Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Araújo, M.B., W. Thuiller, and R.G. Pearson, 2006: Climate warming and the decline of amphibians and reptiles in Europe. Journal of Biogeography, 33, 1712-1728. Araújo, M.B., D. Alagador, M. Cabeza, D. Nogués-Bravo, and W. Thuiller, 2011: Climate change threatens European conservation areas. Ecology Letters, 14, 484-492. Arca, B., G. Pellizzaro, P. Duce, M. Salis, V. Bacciu, D. Spano, A. Ager, and E. Scoccimarro, 2012: Potential changes in fire probability and severity under climate change scenarios in Mediterranean areas. In: Modelling Fire Behaviour and Risk. [Spano, D., V. Bacciu, M. Salis, and C. Sirca (eds.)]. Nuova Stampa Color, Muros, Italy, pp. 92-98. Arctic Climate Impact Assessment, 2005: Arctic Climate Impact Assessment (ACIA). Cambridge University Press, New York, pp. 1042. Armstrong, B.G., Z. Chalab, B. Fenn, S. Hajat, S. Kovats, A. Milojevic, and P. Wilkinson, 2011: Association of mortality with high temperatures in a temperate climate:England and Wales. Journal of Epidemology and Community Health, 65, 340-345. Arnason, R., 2012: Global warming: New challenges for the common fisheries policy? Ocean & Coastal Management, 70, 4-9. Arnell, N., 2011: Incorporating climate change into water resources planning in England and Wales. Journal of the American Water Resources Association, 47(3), 541-549. Artmann, N., D. Gyalistras, H. Manz, and P. Heiselberg, 2008: Impact of climate warming on passive night cooling potential. Building Research & Information, 36(2), 111-128. ARUP, 2008: Your home in a changing climate. Retrofitting existing homes in a changing climate. Report for policy makers. Greater London Authority, London, pp. 75. ARUP, 2011: Adaptation Sub-Committee of the Committee on Climate Change. Analysis of How Land Use Planning Decisions Affect Vulnerability to Climate Risks. Final Report. Ove Arup and Partners Ltd., London, United Kingdom, pp. 132. Arzt, J., W.R. White, B.V. Thomsen, and C.C. Brown, 2010: Agricultural diseases on the move early in the third millennium. Veterinary Pathology, 47(1), 15-27. Aström, D., B. Forsberg, and J. Rocklöv, 2011: Heat wave impact on morbidity and mortality in the elderly population: a review of recent studies. Maturitas, 69(2), 99-105. Aström, D., H. Orru, J. Rocklöv, G. Strandberg, K.L. Ebi, and B. Forsberg, 2013: Heat-related respiratory hospital admissions in Europe in a changing climate: a health impact assesssment. BMJ Open, 3. Avnery, S., D.L. Mauzerall, J. Liu, and L.W. Horowitz, 2011a: Global crop yield reductions due to surface ozone exposure: 1. Year 2000 crop production losses and economic damage. Atmospheric Environment, 45, 2284- 2296. Avnery, S., D.L. Mauzerall, J. Liu, and L.W. Horowitz, 2011b: Global crop yield reductions due to surface ozone exposure: 2. Year 2030 potential crop production losses and economic damage under two scenarios of O3 pollution. Atmospheric Environment, 45, 2297-2309. BACC, 2008: Assessment of Climate Change for the Baltic Sea Basin. Springer, New York, pp. 474. Baccini, M., T. Kosatsky, A. Analitis, H.R. Anderson, M. D'Ovidio, B. Menne, P. Michelozzi, and A. Biggeri, 2011: Impact of heat on mortality in 15 European cities: attributable deaths under different weather scenarios. Journal of Epidemiology and Community Health, 65(1), 64-70. Ballester, J., J. Robine, F.R. Herrmann, and X. Rodo, 2011: Long-term projections and acclimatization scenarios of temperature-related mortality in Europe. Nature Communications, 2(1), Article number358. Bangash, R.F., A. Passuello, M. Sanchez-Canales, M. Terrado, A. Lopez, F.J. Elorza, G. Ziv, V. Acuna, and M. Schuhmacher, 2013: Ecosystem services in Mediterranean river basin: climate change impact on water provisioning and erosion control. The Science of the Total Environment, 458-460, 246-255. Bank of Greece, 2011: The Environmental, Economic and Social Impacts of Climate Change in Greece. Climate Change Impacts Study Committee, Greece, pp. 494. Barredo, J.I., 2010: No upward trend in normalised windstorm losses in Europe: 1970-2008. Natural Hazards and Earth System Sciences, 10(1), 97-104. Barredo, J.I., 2009: Normalised flood losses in Europe: 1970-2006. Natural Hazards and Earth System Sciences, 9(1), 97-104. Barredo, J.I., D. Saurí, and M.C. Llasat, 2012: Assessing trends in insured losses from floods in Spain 1971-2008. Natural Hazards and Earth System Science, 12(5), 1723-1729. Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Barriopedro, D., E.M. Fischer, J. Luterbacher, R.M. Trigo, and R. García-Herrera, 2011: The hot summer of 2010: redrawing the temperature record map of Europe. Science, 332(6026), 220-224. Barstad, I., A. Sorteberg, and M. Dos-Santos, 2012: Present and future offshore wind power potential in Northern Europe based on downscaled global climate runs with adjusted SST and sea ice cover. Renewable Energy, 44, 398-405. Bastian, O., 2013: The role of biodiversity in supporting ecosystem services in Natura 2000 sites. Ecological Indicators, 24, 12-22. Bastola, S., C. Murphy, and J. Sweeney, 2011: The sensitivity of fluvial flood risk in Irish catchments to the range of IPCC AR4 climate change scenarios. Science of the Total Environment, 409(24), 5403-5415. Battaglini, A., G. Barbeau, M. Bindi, and F.W. Badeck, 2009: European winegrowers perceptions of climate change impact and options for adaptation. Regional Environmental Change, 9(2), 61-73. Beaugrand, G., P. Reid, F. Ibanez, J. Lindley, and M. Edwards, 2002: Reorganization of North Atlantic marine copepod biodiversity and climate. Science, 296(5573), 1692-1694. Beaugrand, G. and R.R. Kirby, 2010: Climate, plankton and cod. Global Change Biology, 16(4), 1268-1280. Beaugrand, G., M. Edwards, and L. Legendre, 2010: Marine biodiversity, ecosystem functioning, and carbon cycles. Proceedings of the National Academy of Sciences of the United States of America, 107(22), 10120-10124. Beaugrand, G. and P.C. Reid, 2012: Relationships between North Atlantic salmon, plankton, and hydroclimatic change in the Northeast Atlantic. ICES Journal of Marine Science, 69(9), 1549-1562. Becker, N., 2009: The impact of globalization and climate change on the development of mosquitoes and mosquito- borne diseases in Central Europe [Die Rolle der Globalisierung und Klimaveränderung auf die Entwicklung von Stechmücken und von ihnen übertragenen Krankheiten in Zentral-Europa]. Umweltwissenschaften Und Schadstoff-Forschung, 21(2), 212-222. Belkin, I.M., 2009: Rapid warming of large marine ecosystems. Progress in Oceanography, 81, 207-213. Benavente, D., P. Brimblecombe, and C.M. Grossi, 2008: Salt weathering and climate change. In: New Trends in Analytical, Environmental and Cultural Heritage Chemistry. [Colombini, M.P. and L. Tassi (eds.)]. Transworld Research Network, Kerala, India, pp. 277-286. Beniston, M., 2007: Entering into the 'greenhouse century': recent record temperatures in Switzerland are comparable to the upper temperature quantiles in a greenhouse climate. Geophysical Research Letters, 34(L16710), dpi:10.1029/2007GL030144. Beniston, M., D.B. Stephenson, O.B. Christensen, C.A.T. Ferro, C. Frei, S. Goyette, K. Halsnaes, T. Holt, K. Jylhä, B. Koffi, J. Palutikof, R. Schöll, T. Semmler, and K. Woth, 2007: Future extreme events in European climate: an exploration of regional climate model projections. Climatic Change, 81(Suppl 1), 71-95. Berg, P., C. Moseley, and J.O. Haerter, 2013: Strong increase in convective precipitation in response to higher temperatures. Nature Geoscience, 6(3), 181-185. Bertini, G., T. Amoriello, G. Fabbio, and M. Piovosi, 2011: Forest growth and climate change: Evidences from the ICP-forests intensive monitoring in Italy. Journal of Biogeosciences and Forestry, 4, 262-267. Bett, P.E., H.E. Thornton, and R.T. Clark, 2013: European wind variability over 140 yr . Advances in Science and Research, 10, 51-58. Biesbroek, G.R., R.J. Swart, T.R. Carter, C. Cowan, T. Henrichs, H. Mela, M.D. Morecroft, and D. Rey, 2010a: Europe adapts to climate change: Comparing national adaptation strategies. Global Environmental Change, 20(3), 440-450. Biesbroek, G.R., R.J. Swart, T.R. Carter, C. Cowan, T. Henrichs, H. Mela, M.D. Morecroft, and D. Rey, 2010b: Europe adapts to climate change: Comparing national adaptation strategies. Global Environmental Change, 20(3), 440-450. Bigler, C., O. Bräker, H. Bugmann, M. Dobbertin, and A. Rigling, 2006: Drought as an inciting mortality factor in Scots pine stands of the Valais, Switzerland. Ecosystems, 9(3), 330-343. Bindi, M. and J.E. Olesen, 2011: The responses of agriculture in Europe to climate change. Regional Environmental Change, 11(suppl. 1), 151-158. Bittner, M., E.F. Matthies, D. Dalbokova, and B. Menne, 2013: Are European countries prepared for the next big heat-wave? European Journal of Public Health, 1-5 (doi:10.1093/eurpub/ckt121). Björdal, C.G., 2012: Evaluation of microbial degradation of shipwrecks in the Baltic Sea. International Biodeterioration & Biodegradation, 70(0), 126-140. Blaustein, A.R., S.C. Walls, B.A. Bancroft, J.J. Lawler, C.L. Searle, and S.S. Gervasi, 2010: Direct and indirect effects of climate change on amphibian populations. Diversity, 2(2), 281-313. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bloom, A., V. Kotroni, and K. Lagouvardos, 2008: Climate change impact of wind energy availability in the Eastern Mediterranean using the regional climate model PRECIS. Natural Hazards and Earth System Sciences, 8(6), 1249-1257. Bock, A., T. Sparks, N. Estrella, and A. Menzel, 2011: Changes in the phenology and composition of wine from Franconia, Germany. Climate Research, 50, 69-81. Bolte, A., C. Ammer, M. Löf, P. Madsen, G. Nabuurs, P. Schall, P. Spathelf, and J. Rock, 2009: Adaptive forest management in central Europe: Climate change impacts, strategies and integrative concept. Scandinavian Journal of Forest Research, 24(6), 482-482. Bonazza, A., P. Messina, C. Sabbioni, C.M. Grossi, and P. Brimblecombe, 2009a: Mapping the impact of climate change on surface recession of carbonate buildings in Europe. Science of the Total Environment, 407(6), 2039- 2050. Bonazza, A., C. Sabbioni, P. Messina, C. Guaraldi, and P. De Nuntiis, 2009b: Climate change impact: mapping thermal stress on Carrara marble in Europe. Science of the Total Environment, 407(15), 4506-4512. Bondur, V.G., 2011: Satellite monitoring of wildfires during the anomalous heat wave of 2010 in Russia Izvestiya, Atmospheric and Oceanic Physics, 47(9), 1039-1049. Bormann, H., N. Pinter, and S. Elfert, 2011: Hydrological signatures of flood trends on German rivers: Flood frequencies, flood heights and specific stages. Journal of Hydrology, 404, 50-66. Bosello, F., R.J. Nicholls, J. Richards, R. Roson, and R.S.J. Tol, 2012: Economic impacts of climate change in Europe: Sea-level rise. Climatic Change, 112(1), 63-81. Bottcher, H., P.J. Verkerk, M. Gusti, P. Havllk, and G. Grassi, 2012: Projection of the future EU forest CO2 sink as affected by recent bioenergy policies using two advanced forest management models. GCB Bioenergy, 4(6), 773-783. Botzen, W.J.W. and J.C.J.M. van den Bergh, 2008: Insurance against climate change and flooding in the Netherlands: present, future, and comparison with other countries. Risk Analysis, 28, 413-426. Botzen, W.J.W., J.C.J.H. Aerts, and J.C.J.M. van den Bergh, 2009: Willingness of homeowners to mitigate climate risk through insurance. Ecological Economics, 68, 2265-2277. Botzen, W.J.W., J.C.J.M. van den Bergh, and L.M. Bouwer, 2010a: Climate change and increased risk for the insurance sector: a global perspective and an assessment for the Netherlands. Natural Hazards, 52, 577-598. Botzen, W.J.W., L.M. Bouwer, and J.C.J.M. van den Bergh, 2010b: Climate change and hailstorm damage: Empirical evidence and implications for agriculture and insurance. Resource and Energy Economics, 32(3), 341-362. Bouwer, L.M., J.E. Vermaat, and J.C.J.H. Aerts, 2008: Regional sensitivities of mean and peak river discharge to climate variability in Europe. Journal of Geophysical Research, 113(D19103). Bouwer, L.M., P. Bubeck, and J.C.J.H. Aerts, 2010: Changes in future flood risk due to climate and development in a Dutch polder area. Global Environmental Change, 20(3), 463-471. Boxall, A., A. Hardy, S. Beulke, T. Boucard, L. Burgin, P.D. Falloon, P.M. Haygarth, T. Hutchinson, S. Kovats, G. Leonardi, L.S. Levy, G. Nichols, S.A. Parsons, L. Potts, D. Stone, E. Topp, D.B. Turley, K. Walsh, E.M.H. Wellington, and R.J. Williams, 2009: Impacts of climate change on indirect human exposure to pathogens and chemicals from agriculture: Environmental Health Perspectives, 117(4), 508-514. Bradley, B.A., D.M. Blumenthal, D.S. Wilcove, and L.H. Ziska, 2010: Predicting plant invasions in an era of global change. Trends in Ecology and Evolution, 25(5), 310-318. Branquart, E., K. Verheyen, and J. Latham, 2008: Selection criteria of protected forest areas in Europe: The theory and the real world. Biological Conservation, 11(141), 2795-2806. Breesch, H. and A. Janssens, 2010: Performance evaluation of passive cooling in office buildings based on uncertainty and sensitivity analysis. Solar Energy, 84(8), 1453-1467. Brijs, T., D. Karlis, and G. Wets, 2008: Studying the effect of weather conditions on daily crash counts using a discrete time-series model. Accident Analysis and Prevention, 40(3), 1180-1190. Brimblecombe, P., M.C. Grossi, and I. Harris, 2006: Climate change critical to cultural heritage. In: Heritage Weathering and Conservation. Taylor and Francis, London, UK, pp. 387-393. Brimblecombe, P. and C.M. Grossi, 2008: Millennium-long recession of limestone facades in London. Environmental Geology, 56(3-4), 463-471. Brimblecombe, P. and C.M. Grossi, 2009: Millennium-long damage to building materials in London. Science of the Total Environment, 407(4), 1354-1361. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Brimblecombe, P., 2010a: Climate Change and Cultural Heritage. In: Heritage Climatology . [Lefevre, R.-. and C. Sabbioni (eds.)]. Edipuglia, Bari, Italy, pp. 49-56. Brimblecombe, P., 2010b: Mapping heritage climatologies. In: Effect of Climate Change on Built Heritage . [Bunnik, T., H. de Clercq, R. van Hees, H. Schellen, and L. Schueremans (eds.)]. WTA Publications, Pfaffenhofen, Germany, pp. 18-30. Brimblecombe, P. and C.M. Grossi, 2010: Potential damage to modern building materials from 21st century air pollution. The Scientific World Journal, 10, 116-125. Briner, S., C. Elkin, R. Huber, and A. Gret-Regamy, 2012: Assessing the impacts of economic and climate changes on land-use in mountain regions: A spatial dynamic modelling approach. Agriculture, Ecosystems and Environment, 149, 50-63. Brisson, N., P. Gate, D. Gouache, G. Charmet, F. Oury, and F. Huard, 2010: Why are wheat yields stagnating in Europe? A comprehensive data analysis for France. Field Crops Research, 119(1), 201-212. Brohan, P., J.J. Kennedy, I. Harris, S.F.B. Tett, and P.D. Jones, 2006: Uncertainty estimates in regional and global observed temperature changes: a new data set from 1850. Journal of Geophysical Research, 111(D12106). Brouwer, S., T. Rayner, and D. Huitema, 2013: Mainstreaming climate policy: The case of climate adaptation and the implementation of EU water policy. Environment and Planning C: Government and Policy, 31(1), 134-153. Bryson, J., J. Piper, and M. Rounsevell, 2010: Envisioning futures for climate change policy development: Scenarios use in European environmental policy institutions. Environmental Policy and Governance, 20(5), 283-294. Bubeck, P., H. De Moel, L.M. Bouwer, and J.C.J. H. Aerts, 2011: How reliable are projections of future flood damage? Natural Hazards and Earth System Science, 11(12), 3293-3306. Buestel, D., M. Ropert, J. Prou, and Goulletquer, 2009: History, status and future of oyster culture in France. Journal of Shellfish Research, 28(4), 813-820. Bujosa, A. and J. Roselló, 2012: Climate change and summer mass tourism: the case of Spanish domestic tourism. Climatic Change, 117(1-2), 363-375. Bulkeley, H., 2010: Cities and the Governing of Climate Change. Annual Review of Environment and Resources, 35, 229-253. Busch, G., 2006: Future European agricultural landscapes - What can we learn from existing quantitative land use scenario studies? Agriculture, Ecosystems and Environment, 114(1), 121-140. Butchart, S.H.M., M. Walpole, B. Collen, A. Van Strien, J.P.W. Scharlemann, R.E.A. Almond, J.E.M. Baillie, B. Bomhard, C. Brown, J. Bruno, K.E. Carpenter, G.M. Carr, J. Chanson, A.M. Chenery, J. Csirke, N.C. Davidson, F. Dentener, M. Foster, A. Galli, J.N. Galloway, P. Genovesi, R.D. Gregory, M. Hockings, V. Kapos, J.-. Lamarque, F. Leverington, J. Loh, M.A. McGeoch, L. McRae, A. Minasyan, M.H. Morcillo, T.E.E. Oldfield, D. Pauly, S. Quader, C. Revenga, J.R. Sauer, B. Skolnik, D. Spear, D. Stanwell-Smith, S.N. Stuart, A. Symes, M. Tierney, T.D. Tyrrell, J.-. Vié, and R. Watson, 2010: Global biodiversity: Indicators of recent declines. Science, 328(5982), 1164-1168. Butterworth, M.H., M.A. Semenov, A. Barnes, D. Moran, J.S. West, and B.D.L. Fitt, 2010: North-South divide: contrasting impacts of climate change on crop yields in Scotland and England. Journal of the Royal Society Interface, 7(42), 123-130. Caffarra, A., M. Rinaldi, E. Eccela, V. Rossi, and I. Pertota, 2012: Modelling the impact of climate change on the interaction between grapevine and its pests and pathogens: European grapevine moth and powdery mildew. Agriculture, Ecosystems & Environment, 148, 89-101. Calanca, P., A. Roesch, J. Karsten, and M. Wild, 2006: Global warming and the summertime evapotranspiration regime of the Alpine region. Climatic Change, 79(1-2), 65-78. Callaway, R., A.P. Shinn, S.E. Grenfell, J.E. Bron, G. Burnell, E.J. Cook, M. Crumlish, S. Culloty, K. Davidson, R.P. Ellis, K.J. Flynn, C. Fox, D.M. Green, G.C. Hays, A.D. Hughes, E. Johnston, C.D. Lowe, I. Lupatsch, S. Malham, A.F. Mendzil, T. Nickell, T. Pickerell, A.F. Rowley, M.S. Stanley, D.R. Tocher, J.F. Turnbull, G. Webb, E. Wootton, and R.J. Shields, 2012: Review of climate change impacts on marine aquaculture in the UK and Ireland. Aquatic Conservation: Marine and Freshwater Ecosystems, 22(3), 389-421. Camia, A. and G. Amatulli, 2009: Weather factors and fire danger in the Mediterranean. In: Earth Observation of Wildland Fires in Mediterranean Ecosystems. [Chuvieco, E. (ed.)]. Springer, Berlin, pp. 71-82. Caminade, C., J.M. Medlock, S. leach, K.M. McIntyre, M. Baylis, and A.P. Morse, 2012: Climate suitability of the Asian tiger mosquito Aedes Albopictus in Europe: recent trends and future scenarios. Journal of the Royal Society Interface, 9(75), 2707-2717. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Camps, J.O. and M.C. Ramos, 2012: Grape harvest and yield responses to inter-annual changes in temperature and precipitation in an area of north-east Spain with a Mediterranean climate. International Journal of Biometeorology, 56(5), 853-864. Cantarel, A.M., J.M.G. Bloor, and J. Soussana, 2013: Four years of simulated climate change reduces above-ground productivity and alters functional diversity in a grassland ecosystem. Journal of Vegetation Science, 24(1), 113- 126. Canu, D., C. Solidoro, G. Cossarini, and F. Giorgi, 2010: Effect of global change on bivalve rearing activity and the need for adaptive management. Climate Research, 42, 13-26. Carmichael, C., G. Bickler, R.S. Kovats, D. Pencheon, V. Murray, C. West, and Y. Doyle, 2013: Overheating and hospitals: what do we know? Journal of Hospital Administration, 2(1), DOI: 10.5430/jha.v2n1p1. Carter, J.G., 2011: Climate change adaptation in European cities. Current Opinion in Environmental Sustainability, 3(3), 193-198. Carvalho, A., A. Monteiro, M. Flannigan, S. Solman, A.I. Miranda, and C. Borrego, 2011: Forest fires in a changing climate and their impacts on air quality. Atmospheric Environment, 45(31), 5545-5553. Casalegno, S., G. Amatulli, A. Bastrup-Birk, and T. Houston, 2007: Modelling current and future distribution of European forest categories. Proceedings of Proceedings of the 6th European Conference on Ecological Modelling: Challenges for ecological modelling in a changing world: Global Changes, Sustainability and Ecosystem Based Management, 27-30, November 2007, Trieste, Italy, pp. 1-2. Castebrunet, H., N. Eckert, and G. Giraud, 2012: Snow and weather climatic control on snow avalanche occurrence fluctuations over 50 yr in the French Alps. Climate of the Past, 8(2), 855-875. Castellari, S., 2009: Climate Change, Impacts and Adaptation Strategies in the Alpine Space: Some Results from the INTERREG III B Project ClimChAlp[United Nations Environment Programme (UNEP) Regional Office for Europe Vienna Office (ed.)]. Proceedings of the International Conference Mountains as Early Indicators of Climate Change , 17-18 aprile 2008, Padova, Italy, pp. 81-91. CEA, 2007: Reducing the Social and Economic Impact of Climate Change and Natural Catastrophes: Insurance Solutions and Public-Private Partnerships. European Commission, Brussels, Belgium, pp. 48. CEA, 2009: Tackling Climate Change: The Vital Contribution of Insurers. European Commission, Brussels, Belgium, pp. 1-64. Cellamare, M., M. Leitao, M. Coste, A. Dutartre, and J. Haury, 2010: Tropical phytoplankton taxa in Aquitaine lakes (France). Hydrobiologia, 639(1), 129-145. Charles, E., D. Idier, J. Thiebot, G. Le Cozannet, R. Pedreros, F. Ardhuin, and S. Planton, 2012: Present wave climate in the Bay of Biscay: Spatiotemporal variability and trends from 1958 to 2001. Journal of Climate, 25, 2020-2035. Charlton, M.B. and N.W. Arnell, 2011: Adapting to climate change impacts on water resources in England - an assessment of draft Water Resources Management Plans. Global Environmental Change, 21(1), 238-248. Charru, M., I. Seynave, F. Morneau, and J. Bontemps, 2010: Recent changes in forest productivity: an analysis of national forest inventory data for common beech (Fagus sylvatica L.) in north-eastern France. Forest Ecology and Management, 260, 864-874. Chatterton, J., C. Viavattene, J. Morris, E. Penning-Rowsell, and S. Tapsell, 2010: The Costs of the Summer 2007 Floods in England. Project: SC070039/R1. Environment Agency, Bristol, United Kingdom, pp. 1-41. Chauveau, M., S. Chazot, C. Perrin, P.-. Bourgin, E. Sauquet, J.-. Vidal, N. Rouchy, E. Martin, J. David, T. Norotte, P. Maugis, and X. de Lacaze, 2013: What impacts of climate change on surface hydrology in France by 2070? . La Houille Blanche,(4), 5-15. Cheaib, A., V. Badeau, J. Boe, I. Chuine, C. Delire, E. Dufrene, C. François, E.S. Gritti, M. Legay, C. Pagé, W. Thuiller, N. Viovy, and P. Leadley, 2012: Climate change impacts on tree ranges: model intercomparison facilitates understanding and quantification of uncertainty. Ecology Letters, 15(6), 533-544. Cheung, W.W.L., V.W.Y. Lam, J.L. Sarmiento, K. Kearney, R. Watson, D. Zeller, and D. Pauly, 2010: Large-scale redistribution of maximum fisheries catch potential in the global ocean under climate change. Global Change Biology, 16(1), 24-35. Cheung, W.W.L., J. Pinnegar, G. Merino, M.C. Jones, and M. Barange, 2012: Review of climate change impacts on marine fisheries in the UK and Ireland. Aquatic Conservation: Marine and Freshwater Ecosystems, 22(3), 368- 388. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cheung, W.W.L., J.L. Sarmiento, J. Dunne, T.L. Frölicher, V.W.Y. Lam, M.L.D. Palomares, R. Watson, and D. Pauly, 2013: Shrinking of fishes exacerbates impacts of global ocean changes on marine ecosystems. Nature Climate Change, 3(3), 254-258. Chevalier, V., M. Pepin, L. Plee, and R. Lancelot, 2010: Rift Valley fever - a threat for Europe? Eurosurveillance, 15(10), 18-28. Chiriaco, M.V., L. Perugini, D. Cimini, E. D Amato, R. Valentini, G. Bovio, P. Corona, and A. Barbati, 2013: Comparison of approaches for reporting forest fire-related biomass loss and greenhouse gas emissions in southern Europe. International Journal of Wildland Fire, 22(6), 730-738. Choat, B., S. Jansen, T.J. Brodribb, H. Cochard, S. Delzon, R. Bhaskar, S.J. Bucci, T.S. Feild, S.M. Gleason, U.G. Hacke, A.L. Jacobsen, F. Lens, H. Maherali, J. Martínez-Vilalta, S. Mayr, M. Mencuccini, P.J. Mitchell, A. Nardini, J. Pittermann, R.B. Pratt, J.S. Sperry, M. Westoby, I.J. Wright, and A.E. Zanne, 2012: Global convergence in the vulnerability of forests to drought. Nature, 491(7426), 752-755. Chow, D.H. and G.J. Levermore, 2010: The effects of future climate change on heating and cooling demands in office buildings in the UK. Building Services Engineering Research and Technology, 31(4), 307-323. Christidis, N., G.C. Donaldson, and P.A. Stott, 2010: Causes for the recent changes in cold- and heat-related mortality in England and Wales. Climatic Change, 102(3-4), 539-553. Christierson, B.V., J. Vidal, and S.D. Wade, 2012: Using UKCP09 probabilistic climate information for UK water resource planning. Journal of Hydrology, 424-425, 48-67. Ciais, P., M. Reichstein, N. Viovy, A. Granier, J. Ogee, V. Allard, M. Aubinet, N. Buchmann, C. Bernhofer, A. Carrara, F. Chevallier, N. De Noblet, A.D. Friend, P. Friedlingstein, T. Grunwald, B. Heinesch, P. Keronen, A. Knohl, G. Krinner, D. Loustau, G. Manca, G. Matteucci, F. Miglietta, J.M. Ourcival, D. Papale, K. Pilegaard, S. Rambal, G. Seufert, J.F. Soussana, M.J. Sanz, E.D. Schulze, T. Vesala, and R. Valentini, 2005: Europe-wide reduction in primary productivity caused by the heat and drought in 2003. Nature, 437(7058), 529-533. Ciavola, P., O. Ferreira, P. Haerens, M. Van Koningsveld, and C. Armaroli, 2011: Storm impacts along European coastlines. Part 2: lessons learned from the MICORE project. Environmental Science & Policy, 14(7), 924-933. Ciscar, J.C., A. Iglesias, L. Feyen, L. Szabó, D. Van Regemorter, B. Amelung, R. Nicholls, P. Watkiss, O.B. Christensen, R. Dankers, L. Garrote, C.M. Goodess, A. Hunt, A. Moreno, J. Richards, and A. Soria, 2011: Physical and economic consequences of climate change in Europe. Proceedings of the National Academy of Sciences of the United States of America, 108(7), 2678-2683. Ciscar, J.C., 2009: Climate Change Impacts in Europe. In: Final Report of the PESETA Research Project. JRC Scientific and Technical Reports. European Commission, Seville, Spain, pp. 1-130. Civantos, E., W. Thuiller, L. Maiorano, A. Guisan, and M. Araujo, 2012: Potential Impacts of Climate Change on Ecosystem Services in Europe: The Case of Pest Control by Vertebrates. BioScience, 62, 658-666. Clark, J.M., A. Gallego-Sala, T.E.H. Allott, S. Chapman, T. Farewell, C. Freeman, J.I. House, H.G. Orr, I.C. Prentice, and P. Smith, 2010: Assessing the vulnerability of blanket peat in Great Britain to climate change using an ensemble of statistical bioclimatic envelope models. Climate Research, 45, 131-150. Clark, J.M., H.G. Orr, J. Freer, J.I. House, P. Smith, and C. Freeman, 2010a: Assessment of projected changes in upland environments using simple climatic indexes. Climate Research, 45, 87-104. Clemo, K., 2008: Preparing for climate change: Insurance and small business. Geneva Papers on Risk and Insurance: Issues and Practice, 33(1), 110-116. Cogan, D.G., 2008: Corporate governance and climate change: the banking sector. In: Ceres Report. Ceres, Boston, USA, pp. 1-64. Conraths, F.J. and T.C. Mettenleiter, 2011: Globalisation and change of climate: Growing risk for livestock epidemics in Germany [Globalisierung und Klimawandel: Steigendes Risiko für Tierseuchen in Deutschland]. Zuchtungskunde, 83(1), 21-26. Corobov, R., S. Sheridan, N. Opopol, and K. Ebi, 2012: Heat-related mortality in Moldova: the summer of 2007. International Journal of Climatology, 33(11), 2551-2560. Corobov, R., S. Sheridan, K. Ebi, and N. Opopol, 2013: Warm season temperature-mortality relationships in Chisinau (Moldova). International Journal of Atmospheric Sciences, 2013(Article ID 346024), 1-9 (doi:10.1155/2013/346024). Corti, T., V. Muccione, P. Kollner-Heck, D. Bresch, and S.I. Seneviratne, 2009: Simulating past droughts and associated building damages in France. Hydrology and Earth System Sciences, 13(9), 1739-1747. Coumou, D. and S. Rahmstorf, 2012: A decade of weather extremes. Nature Climate Change, 2(7), 491-496. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Crescio, M.I., F. Forastiere, C. Maurella, F. Ingravalle, and G. Ru, 2010: Heat-related mortality in dairy cattle: A case crossover study. Preventative Veterinary Medicine, 97, 191-197. Crichton, D., 2006: Climate change and its effects on small business in the UK . AXA Insurance UK, UK, pp. 1-46. Crichton, D., 2007: The Hull floods of June 2007. Some insurance industry implications. [Benfield UCL Hazard Research Centre, UCL, London (ed.)]. Crook, J.A., L.A. Jones, P.M. Forster, and R. Crook, 2011: Climate change impacts on future photovoltaic and concentrated solar power energy output. Energy & Environmental Science, 4(9), 3101-3109. Crump, D., A. Dengel, and M. Swainson, 2009: Indoor Air Quality in Highly Energy Efficient Homes - A Review. (NF18). NHBC Foundation, IHS BRE Press, Milton Keynes, United Kingdom, pp. 1-84. Daccache, A. and N. Lamaddalena, 2010: Climate change impacts on pressurised irrigation systems. Proceedings of the Institution of Civil Engineers-Engineering Sustainability, 163(2), 97-105. Daccache, A., C. Keay, R.J.A. Jones, E.K. Weatherhead, M.A. Stalham, and J.W. Knox, 2012: Climate change and land suitability for potato production in England and Wales: Impacts and adaptation. Journal of Agricultural Science, 150(2), 161-177. Damigos, D., 2012: Monetizing the impacts of climate change on the Greek mining sector. Mitigation and Adaptation Strategies for Global Change, 17(8), 865-878. Dammers, E., 2010: Making territorial scenarios for Europe. Futures, 42, 785-793. Dankers, R., O.B. Christensen, L. Feyen, M. Kalas, and A. de Roo, 2007: Evaluation of very high-resolution climate model data for simulating flood hazards in the Upper Danube Basin. Journal of Hydrology, 347(3-4), 319-331. Dankers, R. and L. Feyen, 2008: Climate change impact on flood hazard in Europe: An assessment based on high- resolution climate simulations. Journal of Geophysical Research, 113, D19105. Daufresne, M., P. Bady, and J.F. Fruget, 2007: Impacts of global changes and extreme hydroclimatic events on macroinvertebrate community structures in the French Rhône River. Oecologia, 151(3), 544-559. Daufresne, M., K. Lengfellner, and U. Sommer, 2009: Global warming benefits the small in aquatic ecosystems. Proceedings of the National Academy of Sciences of the United States of America, 106(31), 12788-93. Davies, M. and T. Oreszczyn, 2012: The unintended consequences of decarbonising the built environment: A UK case study . Energy and Buildings, 46, 80-85. Davoudi, S., M. Wishardt, and I. Strange, 2010: The ageing of Europe: Demographic scenarios of Europe s futures . Futures, 42, 794-803. Dawson, R.J., T. Ball, J. Werritty, A. Werritty, J.W. Hall, and N. Roche, 2011: Assessing the effectiveness of non- structural ood management measures in the Thames Estuary under conditions of socio-economic and environmental change. Global Environmental Change, 21, 628-646. Day, A.R., P.G. Jones, and G.G. Maidment, 2009: Forecasting future cooling demand in London. Energy and Buildings, 41(9), 942-948. Day, J.W., R.R. Christian, D.M. Boesch, A. Yánez-Arancibia, J. Morris, R.R. Twilley, L. Naylor, L. Schaffner, and C. Stevenson, 2008: Consequences of climate change on the ecogeomorphology of coastal wetlands. Estuaries and Coasts, 31(3), 477-491. DCLG, 2012: Investigation into overheating in homes: literature review. UK Department of Communities and Local Government, London, pp. 124. De Freitas, C.R., D. Scott, and G. McBoyle, 2008: A second generation climate index for tourism (CIT): specification and verification. International Journal of Biometeorology, 52(5), 399-407. de Graaff, M., C. Van Kessel, and J. Six, 2009: Rhizodeposition-induced decomposition increases N availability to wild and cultivated wheat genotypes under elevated CO2. Soil Biology & Biochemistry, 41(6), 1094-1103. de Moel, H., J. van Alphen, and J.C.J.H. Aerts, 2009: Flood maps in Europe methods, availability and use. Natural Hazards and Earth System Sciences, 9(2), 289-301. De Wit, M., M. Londo, and A. Faaij, 2011: Productivity developments in European agriculture: Relations to and opportunities for biomass production. Renewable and Sustainable Energy Reviews, 15(5), 2397-2412. Debernard, J.B. and L.P. Ryed, 2008: Future wind, wave and storm surge climate in the Northern Seas: a revisit. Tellus A, 60(3), 427-438. del Barrio, G., P.A. Harrison, P.M. Berry, N. Butt, M.E. Sanjuan, R.G. Pearson, and T. Dawson, 2006: Integrating multiple modelling approaches to predict the potential impacts of climate change on species' distributions in contrasting regions: Comparison and implications for policy. Environmental Science and Policy, 9(2), 129-147. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Della Bella, V., M. Bazzanti, M.G. Dowgiallo, and M. Iberite, 2008: Macrophyte diversity and physico-chemical characteristics of Tyrrhenian coast ponds in central Italy: Implications for conservation. Hydrobiologia, 597(1), 85-95. Dell'Aquila, A., S. Calmanti, P.M. Ruti, M.V. Struglia, G. Pisacane, A. Carillo, and G. Sannino, 2012: Impacts of seasonal cycle fluctuations over the Euro-Mediterranean area using a Regional. Climate Research, 52, 135-157. Delpla, I., E. Baures, A.-. Jung, and O. Thomas, 2011: Impacts of rainfall events on runoff water quality in an agricultural environment in temperate areas. Science of the Total Environment, 409, 1683-1688. Delta Committee, 2008: Working Together With Water. A Living Land Builds for its Future. [ten Brinke, W.B.M. (ed.)]. Secretariat Delta Committee, Netherlands, pp. 1-26. Dessai, S. and M. Hulme, 2007: Assessing the robustness of adaptation decisions to climate change uncertainties: A case study on water resources management in the East of England. Global Environmental Change, 17(1), 59-72. Devictor, V., R. Julliard, D. Couvet, and F. Jiguet, 2008: Birds are tracking climate warming, but not fast enough. Proceedings of the Royal Society B: Biological Sciences, 275, 2743-2748. D'Ippoliti, D., P. Michelozzi, C. Marino, F. de'Donato, B. Menne, K. Katsouyanni, U. Kirchmayer, A. Analitis, M. Medina-Ramon, A. Paldy, R. Atkinson, S. Kovats, L. Bisanti, A. Schneider, A. Lefranc, C. Iniguez, and C.A. Perucci, 2010: The impact of heat waves on mortality in 9 European cities: results from the EuroHEAT project. Environmental Health: A Global Access Science Source, 9, 37 (doi:10.1186/1476-069X-9-37). Dixon, N. and E. Brook, 2007: Impact of predicted climate change on landslide reactivation: case study of Mam Tor. UK Landslides, 4, 137-147. Dobney, K., C.J. Baker, L. Chapman, and A.D. Quinn, 2010: The future cost to the United Kingdom's railway network of heat-related delays and buckles caused by the predicted increase in high summer temperatures owing to climate change. Proceedings of the Institution of Mechanical Engineers, Part F. Journal of Rail and Rapid Transit, 224(1), 25-34. Dolinar, M., B. Vidrih, L. Kajfe¾-Bogataj, and S. Medvec, 2010: Predicted changes in energy demands for heating and cooling due to climate change. Physics and Chemistry of the Earth, 35(1-2), 100-106. Donat, M.G., G.C. Leckebusch, J.G. Pinto, and U. Ulbrich, 2010: European storminess and associated circulation weather types: future changes deduced from a multi-model ensemble of GCM simulations. Climate Research, 42(1), 27-43. Donat, M.G., G.C. Leckebusch, S. Wild, and U. Ulbrich, 2011: Future changes in European winter storm losses and extreme wind speeds inferred from GCM and RCM multi-model simulations. Natural Hazards and Earth System Sciences, 11(5), 1351-1370. Donatelli, M., A.K. Srivastava, G. Duveiller, and S. Niemeyer, 2012: Estimating Impact Assessment and Adaptation Strategies under Climate Change Scenarios for Crops at EU27 Scale. In: Environmental Modelling and Software [R. Seppelt, R., A.A. Voinov, S. Lange, D. Bankamp(eds.)]. Proceedings of Managing Resources of a Limited Planet, Sixth Biennial Meeting, July 2012, Leipzig, Germany, pp. 1-8. Doney, S.C., M. Ruckelshaus, J. Emmett Duffy, J.P. Barry, F. Chan, C.A. English, H.M. Galindo, J.M. Grebmeier, A.B. Hollowed, N. Knowlton, J. Polovina, N.N. Rabalais, W.J. Sydeman, and L.D. Talley, 2011: Climate Change impacts on marine ecosystems. Annual Review of Marine Science, 4, 11-37. Drenkhan, R., T. Kurkela, and M. Hanso, 2006: The relationship between the needle age and the growth rate in Scots pine (Pinus sylvestris): a retrospective analysis by needle trace method (NTM). European Journal of Forest Research, 125, 397-405. Dronin, N. and A. Kirilenko, 2011: Climate change, food stress, and security in Russia. Regional Environmental Change, 11(SUPPL. 1), 167-178. Duarte Alonso, A. and M.A. O Neill, 2011: Climate change from the perspective of Spanish wine growers: a three- region study. British Food Journal, 113(2), 205-221. Ducharne, A., F. Habets, C. Pagé, E. Sauquet, P. Viennot, M. Déqué, S. Gascoin, A. Hachour, E. Martin, L. Oudin, L. Terray, and D. Thiéry, 2010: Climate change impacts on water resources and hydrological extremes in northern France[Carrera, J. (ed.)]. Proceedings of XVIII International Conference on Computational Methods in Water Resources, CIMNE, 21-24, June 2010, Barcelona, Spain, pp. 1-8. Ducharne, A., C. Baubion, N. Beaudoin, M. Benoit, G. Billen, N. Brisson, J. Garnier, H. Kieken, S. Lebonvallet, E. Ledoux, B. Mary, C. Mignolet, X. Poux, E. Sauboua, C. Schott, S. Thery, and P. Viennot, 2007: Long term prospective of the Seine River system: Confronting climatic and direct anthropogenic changes. Science of the Total Environment, 375(1-3), 292-311. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Duchene, E., F. Huard, V. Dumas, C. Schneider, and D. Merdinoglu, 2010: The challenge of adapting grapevine varieties to climate change. Climate Research, 41(3), 193-204. Dullinger, S., A. Gattringer, W. Thuiller, D. Moser, N.E. Zimmermann, A. Guisan, W. Willner, C. Plutzar, M. Leitner, T. Mang, M. Caccianiga, T. Dirnböck, S. Ertl, A. Fischer, J. Lenoir, J. Svenning, A. Psomas, D.R. Schmatz, U. Silc, P. Vittoz, and K. Hülber, 2012: Extinction debt of high-mountain plants under twenty-first- century climate change. Nature Climate Change, 2(8), 619-622. Dumollard, G. and A. Leseur, 2011: Drawing Up A National Adaptation Policy: Feedback on Five European Case Studies. In: Climate Report. Research on the Economics of Climate Change. CDC Climat Research, Paris, France, pp. 1-32. Durant, J.M., D.O. Hjermann, G. Ottersen, and N.C. Stenseth, 2007: Climate and the match or mismatch between predator requirements and resource availability. Climate Research, 33(3), 271-283. Dury, M., A. Hambuckers, P. Warnant, A. Henrot, E. Favre, M. Ouberdous, and L. François, 2011: Responses of European forest ecosystems to 21st century climate: assessing changes in interannual variability and fire intensity. IForest, 4, 82-99. EA, 2011: TE2100 Strategic Outline Programme. Environment Agency, London. Easterling, W.E., P.K. Aggarwal, P. Batima, K.M. Brander, L. Erda, S.M. Howden, A. Kirilenko, J. Morton, J. Soussana, J. Schmidhuber, and F.N. Tubiello, 2007: Chapter 5: Food, fibre and forest products. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, United Kingdom, pp. 273-313. EC, 2009a: White Paper: Adapting to Climate Change: Towards a European Framework for Action. European Commission, Brussels, Belgium, pp. 1-16. EC, 2009b: River Basin Management in a Changing Climate. Guidance Document No. 24. In: Common Implementation Strategy for the Water Framework Directive (2000/60/EC). European Commission, Brussels, Belgium, pp. 1-132. EC, 2011: Our Life Insurance, Our Natural Capital: An EU Biodiversity Strategy to 2020. COM(2011) 244 Final. In: Communication from the Commission to the European Parliament, The Council, The Economic and Social Committee and The Committee of the Regions. European Commission, Brussels, Belgium, pp. 1-17. EC, 2013a: Communication from the Commission to the European Parliament, The Council, The European Economic and Social Committee and the Committee of the Regions. An EU Strategy on adaptation to climate change. COM(2013) 216 Final. European Commission, Brussels, pp. 1-11. EC, 2013b: Climate-Adapt. European Climate Adaptation Platform. http://climate-adapt.eea.europa.eu/ European Commission, Brussels. ECDC, 2009: Technical Report: Development of Aedes albopictus Risk Map. European Centre for Disease Prevention and Control, Stockholm, Sweden, pp. 1-45. ECDC, 2012: Technical Report: The Climatic Suitability for Dengue Transmission in Continental Europe. European Centre for Disease Prevention and Control, Stockholm, Sweden, pp. 1-22. ECHOES Country report, 2009: COST action FP0703. [Cacot, E. and Peyron, J. (eds.)], pp. 1-41. Eckert, N., E. Parent, R. Kies, and H. Baya, 2010: A spatio-temporal modelling framework for assessing the fluctuations of avalanche occurrence resulting from climate change: Application to 60 years of data in the Northern French Alps. Climatic Change, 101(3), 515-553. Edwards, M. and A.J. Richardson, 2004: Impact of climate change on marine pelagic phenology and trophic mismatch. Nature, 430(7002), 881-884. EEA, 2009: Water Resources Across Europe Confronting Water Scarcity and Drought. European Environment Agency, Copenhagen, Denmark, pp. 55. EEA, 2010a: The European Environment, State and Outlook 2010. Water Resources: Quantity and Flows. European Environment Agency, Copenhagen, Denmark, pp. 32. EEA, 2010b: Tracking Progress Towards Kyoto and 2020 Targets in Europe. European Environment Agency, Copenhagen, Denmark, pp. 107. EEA, 2010c: Mapping the Impacts of Natural Hazards and Technological Accidents in Europe: An Overview of the Last Decade. European Environment Agency, Copenhagen, Denmark, pp. 144. EEA, 2010d: 10 messages for 2010 Marine Ecosystems. European Environment Agency, Copenhagen, Denmark, pp. 16. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 EEA, 2012: Climate Change, Impacts and Vulnerability in Europe 2012, an Indicator-Based Report. European Environment Agency, Copenhagen, Denmark, pp. 1-304. EEA, 2013: Adaptation in Europe - Addressing Risks and Opportunities From Climate Change in the Context of Socio-Economic Developments. European Environment Agency, Copenhagen, Denmark, pp. 1-136. Ellwanger, G., A. Ssymank, A. Paulsch, and C. Paulsch, 2011: Natura 2000 and Climate Change - State of knowledge: First results of an international workshop on the Isle of Vilm. Natur Und Landschaft, 86(1), 15-18. ELME, 2007: European Lifestyles and Marine Ecosystems: Exploring Challenges for Managing Europe s Seas. [Langmead, O., McQuatters-Gollop, A. and Mee, L.D. (eds.)]. University of Plymouth Marine Institute, Plymouth, United Kingdom, pp. 1-43. Elzinga, J.A., S. van Nouhuys, D. van Leeuwen, and A. Biere, 2007: Distribution and colonisation ability of three parasitoids and their herbivorous host in a fragmented landscape. Basic and Applied Ecology, 8(1), 75-88. Endler, C., K. Oehler, and A. Matzarakis, 2010: Vertical gradient of climate change and climate tourism conditions in the Black Forest. International Journal of Biometeorology, 54(1), 45-61. Endler, C. and A. Matzarakis, 2011b: Climatic potential for tourism in The Black Forest, Germany - winter season. International Journal of Biometeorology, 55(3), 339-351. Endler, C. and A. Matzarakis, 2011a: Climatic potential for tourism in the black forest,Germany-winter season. International Journal of Biometeorology, 55(3), 339-351. Engelhard, G.H., J.R. Ellis, M.R. Payne, R. Ter Hofstede, and J.K. Pinnegar, 2011: Ecotypes as a concept for exploring responses to climate change in fish assemblages. ICES Journal of Marine Science, 68(3), 580-591. Engler, R., C. Randin, W. Thuiller, S. Dullinger, N.E. Zimmermann, M.B. Araújo, P.B. Pearman, C.H. Albert, P. Choler, X. de Lamo, T. Dirnböck, D. Gómez-García, J. Grytnes, E. Heegard, F. Histad, G. Le Lay, D. Nogues- Bravo, S. Normand, C. Piédalu, M. Puscas, M. Sebastia, A. Stanisci, J. Theurillat, M. Trivedi, P. Vittoz, and A. Guisan, 2011: 21st century climate change threatens mountain flora unequally across Europe. Global Change Biology, 17, 2330-2341. Environmental Agency, 2009: Thames Estuary 2100, Consultation Document Environmental Agency, UK. Environmental Agency, United Kingdom, pp. 214. Eskeland, G.S. and T.K. Mideksa, 2010: Electricity demand in a changing climate. Mitigation and Adaptation Strategies for Global Change, 15(8), 877-897. ESPACE, 2007: European Spatial Planning Adapting to Climate Events - Final Report. The Environment Department, Hampshire County Council, The Castle, Winchester SO23 8UD, United Kingdom, pp. 8. Eugenio-Martin, J.L. and J.A. Campos-Soria, 2010: Climate in the region of origin and destination choice in outbound tourism demand. Tourism Management, 31(6), 744-753. European Commission, 2013: Principles and recommendations for integrating climate change adaptation considerations under the 2014-2020 European Maritime and Fisheries Fund operational programmes. In: Brussels, 30.7.2013. SWD(2013) 299 final. European Commission, Brussels. European Parliament and Council, 2007: Directive of the European Parliament and of the Council of 23 October 2007 on the assessment and management of flood risks 2007/60/EC, pp. L 288/27-L 288/34. Eurostat, 2009: Forestry Statistics. European Union, Luxembourg, pp. 1-172. Eurostat, 2011a: Migrants in Europe. A Statistical Portrait of the First and Second Generation. European Union, Luxembourg, pp. 1-155. Eurostat, 2011b: Labour Market Statistics. European Union, Luxembourg, pp. 1-109. Falloon, P. and R. Betts, 2010: Climate impacts on European agriculture and water management in the context of adaptation and mitigation-The importance of an integrated approach. Science of the Total Environment, 408(23), 5667-5687. FAO, 2008a: Climate change: Implications for Food Safety. Food and Agriculture Organization, Rome, Italy, pp. 49. FAO, 2008b: Climate change impacts on forest health. Food and Agriculture Organization, Rome, Italy, pp. 38. Feehan, J., M. Harley, and J. Van Minnen, 2009: Climate change in Europe. 1. Impact on terrestrial ecosystems and biodiversity. A review. Agronomy for Sustainable Development, 29(3), 409-421. Fernandes, P.M., A. Luz, and C. Loureiro, 2010: Changes in wildfire severity from maritime pine woodland to contiguous forest types in the mountains of northwestern Portugal. Forest Ecology and Management, 260(5), 883-892. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ferrara, R.M., P. Trevisiol, M. Acutis, G. Rana, G.M. Richter, and N. Baggaley, 2010: Topographic impacts on wheat yields under climate change: two contrasted case studies in Europe. Theoretical and Applied Climatology, 99(1-2), 53-65. Ferron, C., D. Trewick, P. Le Conte, E.R. Batard, and L. Girard, 2006: Heat stroke in hospital patients during the summer 2003 heat wave: a nosocomial disease. Presse Medicale, 25(2), 196-199. Feyen, L., R. Dankers, K. Bódis, P. Salamon, and J.I. Barredo, 2012: Fluvial flood risk in Europe in present and future climates. Climatic Change, 112, 47-62. Feyen, L., J.I. Barredo, and R. Dankers, 2009: Implications of global warming and urban land use change on flooding in Europe. In: Water & Urban Development Paradigms - Towards an Integration of Engineering, Design and Management Approaches [Feyen, J., K. Shannon, M. Neville(eds.)], pp. 217-225. Feyen, L. and R. Dankers, 2009: Impact of global warming on streamflow drought in Europe. Journal of Geophysical Research: Atmospheres, 114(17), Article numberD17116. Filz, K.J., J.O. Engler, J. Stoffels, M. Weitzel, and T. Schmitt, 2013: Missing the target? A critical view on butterfly conservation efforts on calcareous grasslands in south-western Germany. Biodiversity and Conservation, 22(10), 2223-2241. Finger, R., W. Hediger, and S. Schmid, 2011: Irrigation as adaptation strategy to climate change-a biophysical and economic appraisal for Swiss maize production. Climatic Change, 105(3-4), 509-528. Fischer, D., S.M. Thomas, and C. Beierkuhnlein, 2010a: Temperature-derived potential for the establishment of phlebotomine sandflies and visceral leishmaniasis in Germany. Geospatial Health, 5(1), 59-69. Fischer, G., S. Prieler, H. van Velthuizen, G. Berndes, A. Faaij, M. Londo, and M. de Wit, 2010b: Biofuel production potentials in Europe: Sustainable use of cultivated land and pastures, Part II: Land use scenarios. Biomass and Bioenergy, 34(2), 173-187. Fischer, L., R. Purves, C. Huggel, J. Noetzli, and W. Haeberli, 2012: On the influence of topographic, geological and cryospheric factors on rock avalanches and rockfalls in high-mountain areas. Natural Hazards and Earth System Science, 12(1), 241-254. Fisher, D., C. Thomas, F. Niemitz, and C. Reineking, 2011: Projection of Climate suitability for Aedes albopictus Skuse (Culicidae) in Europe under climate change conditions. Global and Planetary Change, 78, 54-65. Forkel, R. and R. Knoche, 2006: Regional climate change and its impact on photo-oxidant concentrations in sourthern Germany: simulations with a coupled regional climate-chemistry model. Journal of Geophysical Research Atmospheres, 111(D12) (DOI: 10.1029/2005JD006748). Forkel, R. and R. Knoche, 2007: Nested regional climate-chemistry simulations for central Europe. Comptes Rendus Geoscience, 339(11-12), 734-746. Forsius, M., S. Anttila, L. Arvola, I. Bergström, H. Hakola, H.I. Heikkinen, J. Helenius, Hyvärinen, M., Jylhä, K., J. Karjalainen, T. Keskinen, K. Laine, E. Nikinmaa, P. Peltonen- Sainio, M. Pulkkanen, K. Rankinen, M. Reinikainen, H. Setälä, and J. Vuorenmaa, 2013: Impacts and adaptation options of climate change on ecosystem services in Finland: a model-based study. Current Opinion in Environmental Sustainability, 5(1), 26- 40. Förster, H. and J. Lilliestam, 2010: Modeling thermoelectric power generation in view of climate change. Regional Environmental Change, 10(4), 327-338. Founda, D. and C. Giannakopoulos, 2009: The exceptionally hot summer of 2007 in Athens, Greece - A typical summer in the future climate? Global and Planetary Change, 67(3-4), 227-236. Fronzek, S., M. Luoto, and T.R. Carter, 2006: Potential effect of climate change on the distribution of palsa mires in subarctic Fennoscandia. Climate Research, 32(1), 1-12. Fronzek, S., T.R. Carter, J. Räisänen, L. Ruokolainen, and M. Luoto, 2010: Applying probabilistic projections of climate change with impact models: A case study for sub-arctic palsa mires in Fennoscandia. Climatic Change, 99(3), 515-534. Fronzek, S., T.R. Carter, and M. Luoto, 2011: Evaluating sources of uncertainty in modelling the impact of probabilistic climate change on sub-arctic palsa mires. Natural Hazards and Earth System Science, 11(11), 2981-2995. Fronzek, S., T.R. Carter, and K. Jylhä, 2012: Representing two centuries of past and future climate for assessing risks to biodiversity in Europe. Global Ecology and Biogeography, 21(1), 19-35. Fuhrer, J., M. Beniston, A. Fischlin, C. Frei, S. Goyette, K. Jasper, and C. Pfister, 2006: Climate risks and their impact on agriculture and forests in Switzerland. Climatic Change, 79(1-2), 79-102. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Fuhrer, J., 2009: Ozone risk for crops and pastures in present and future climates. Naturwissenschaften, 96(2), 173- 194. Fujihara, Y., K. Tanaka, T. Watanabe, T. Nagano, and T. Kojiri, 2008: Assessing the impacts of climate change on the water resources of the Seyhan River Basin in Turkey: Use of dynamically downscaled data for hydrologic simulations. Journal of Hydrology, 353(1-2), 33-48. Furrer, B., V. Hoffmann, and M. Swoboda, 2009: Banking & Climate Change: Opportunities and Risks. SAM, ETH, and ZHAW., Zurich, Switzerland, pp. 1-51. Gale, P., B. Stephenson, A. Brouwer, M. Martinez, A. de la Torre, J. Bosch, M. Foley-Fisher, P. Bonilauri, A. Lindström, R.G. Ulrich, C.J. de Vos, M. Scremin, Z. Liu, L. Kelly, and M.J. Munoz, 2012: Impact of climate change on risk of incursion of Crimean-Congo haemorrhagic fever virus in livestock in Europe through migratory birds. Journal of Applied Microbiology, 112(2), 246-257. Gallego-Sala, A.V., J.M. Clark, J.I. House, H.G. Orr, I.C. Prentice, P. Smith, T. Farewell, and S.J. Chapman, 2010: Bioclimatic envelope model of climate change impacts on blanket peatland distribution in Great Britain. Climate Research, 45, 151-162. Gao, X. and F. Giorgi, 2008: Increased aridity in the Mediterranean region under greenhouse gas forcing estimated from high resolution simulations with a regional climate model. Global and Planetary Change, 62(3-4), 195- 209. Garcia-Fayos, P. and E. Bochet, 2009: Indication of antagonistic interaction between climate change and erosion on plant species richness and soil properties in semiarid Mediterranean ecosystems. Global Change Biology, 15(2), 306-318. García-López J.M. and C. Alluéa, 2011: Modelling phytoclimatic versatility as a large scale indicator of adaptive capacity to climate change in forest ecosystems. Ecological Modelling, 222(8), 1436-1447. García-Ruiz, J.M., J.I. López-Moreno, S.M. Vicente-Serrano, T. Lasanta-Martínez, and S. Baguería, 2011: Mediterranean water resources in a global change scenario. Earth-Science Reviews, 105(3-4), 121-139. Gardiner, B., K. Blennow, J. Carnus, P. Fleischer, F. Ingemarson, G. Landmann, M. Lindner, M. Marzano, B. Nicoll, C. Orazio, J. Peyron, M. Reviron, M. Schelhaas, A. Schuck, M. Spielmann, and T. Usbeck, 2010: Destructive Storms in European Forests: Past and Forthcoming Impacts. Final report to European Commission - DG Environment. European Forest Institute, Atlantic European Regional Office - EFIATLANTIC, Bordeaux, France, pp. 138. Garza-Gil, M., J. Torralba-Cano, and M. Varela-Lafuente, 2010: Evaluating the economic effects of climate change on the European sardine fishery. Regional Environmental Change, 11(1), 87-95. Gaslikova, L., A. Schwerzmann, C.C. Raible, and T.F. Stocker, 2011: Future storm surge impacts on insurable losses for the North Sea region. Natural Hazards and Earth System Sciences, 11(4), 1205-1216. Gehrig-Fasel, J., A. Guisan, and N.E. Zimmermann, 2007: Tree line shifts in the Swiss Alps: Climate change or land abandonment? Journal of Vegetation Science, 18(4), 571-582. GIA, 2011: The Climate Change Challenge: Answers and Demands of German Insurers. German Insurance Association, Berlin, Germany, pp. 1-20. Giannakopoulos, C., P. Le Sager, M. Bindi, M. Moriondo, E. Kostopoulou, and C.M. Goodess, 2009: Climatic changes and associated impacts in the Mediterranean resulting from a 2 oC global warming. Global and Planetary Change, 68(3), 209-224. Giannakopoulos, C., E. Kostopoulou, K.V. Varotsos, K. Tziotziou, and A. Plitharas, 2011: An integrated assessment of climate change impacts for Greece in the near future. Regional Environmental Change, 11(4), 829-843. Gifford, R., L. Steg, and J.P. Reser, 2011: Environmental Pyschology. In: The IAAP Handbook of Applied Psychology. [Martin, P.R., M.C. Cheung, L. Kyrios, M. Littlefield, J.B. Knowles, M. Overmier et al. (eds.)]. Wiley-Blackwell, Chichester, pp. 440-471. Gilgen, A.K., C. Signarbieux, U. Feller, and N. Buchmann, 2010a: Competitive advantage of Rumex obtusifolius L. might increase in intensively managed temperate grasslands under drier climate. Agriculture, Ecosystems & Environment, 135(1-2), 15-23. Gilgen, A.K., C. Signarbieux, U. Feller, and N. Buchmann, 2010b: Competitive advantage of Rumex obtusifolius L. might increase in intensively managed temperate grasslands under drier climate. Agriculture Ecosystems & Environment, 135(1-2), 15-23. Gill, S., J. Handley, R. Ennos, and S. Pauleit, 2007: Adapting cities for climate change: the role of the green infrastructure. Built Environment, 33(1), 115-133. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Giuggiola, A., T.M. Kuster, and S. Saha, 2010: Drought-induced mortality of Scots pines at the southern limits of its distribution in Europe: causes and consequences. Journal of Biogeosciences and Forestry, 3, 95-97. Giuntoli, I., B. Renard, J.-. Vidal, and A. Bard, 2013: Low flows in France and their relationship to large-scale climate indices. Journal of Hydrology, 482, 105-118. GLA,, 2010: The Draft Climate Change Adaptation Strategy for London, Public Consultation Draft . Greater London Authority, London, United Kingdom, pp. 1-138. Glenk, K. and A. Fisher, 2010: Insurance, prevention or just wait and see? Public preferences for water management strategies in the context of climate change. Ecological Economics, 69, 2279-2291. Goderniaux, P., S. Brouyére, S. Blenkinsop, A. Burton, H.J. Fowler, P. Orban, and A. Dassargues, 2011: Modeling climate change impacts on groundwater resources using transient stochastic climatic scenarios. Water Resources Research, 47(12), Article num. W12516. Golombek, R., S. Kittlesen, and I. Haddeland, 2012: Climate change: impacts on electricity markets in Western Europe. Climatic Change, 113, 357-370. Gómez-Rodríguez, C., J. Bustamante, and C. Díaz-Paniagua, 2010: Evidence of hydroperiod shortening in a preserved system of temporary ponds. Remote Sensing, 2(6), 1439-1462. Gonzalez-Camacho, J., J.C. Mailhol, and F. Ruget, 2008: Local impact of increasing CO2 in the atmosphere on maize crop water productivity in the Drome valley, France. Irrigation and Drainage, 57(2), 229-243. Goode, J., 2012: Viticulture: Fruity with a hint of drought. Nature, 492(7429), 351-353. Goodess, C., D. Jacob, M. Déqué, J. Guttiérrez, R. Huth, E. Kendon, G. Leckebusch, P. Lorenz, and V. Pavan, 2009: Downscaling methods, data and tools for input to impacts assessments. In: ENSEMBLES: Climate Change and its Impacts: Summary of Research and Results from the ENSEMBLES Project. [van der Linden, P. and J.F.B. Mitchell (eds.)]. Met Office Hadley Centre, Exeter, United Kingdom, pp. 59-78. Görgen, K., J. Beersma, H. Buiteveld, G. Brahmer, M. Carambia, O.d. Keizer, P. Krahe, E. Nilson, R. Lammersen, C. Perrin, and D. Volken, 2010: Assessment of Climate Change Impacts on Discharge in the River Rhine Basin. Results of the RheinBlick2050 project. International Commission for the Hydrology of the Rhine Basin, Lelystad, pp. 228. Gottfried, M., H. Pauli, A. Futschik, M. Akhalkatsi, P. Barancok, J.L. Benito Alonso, G. Coldea, J. Dick, B. Erschbamer, M.R. Fernández Calzado, G. Kazakis, J. Krajci, P. Larsson, M. Mallaun, O. Michelsen, D. Moiseev, P. Moiseev, U. Molau, A. Merzouki, L. Nagy, G. Nakhutsrishvili, B. Pedersen, G. Pelino, M. Puscas, G. Rossi, A. Stanisci, J.-. Theurillat, M. Tomaselli, L. Villar, P. Vittoz, I. Vogiatzakis, and G. Grabherr, 2012: Continent-wide response of mountain vegetation to climate change. Nature Climate Change, 2(2), 111-115. Graux, A., R. Lardy, G. Bellocchi, and J. Soussana, 2012: Global warming potential of French grassland-based dairy livestock systems under climate change. Regional Environmental Change, 12(4), 751-763. Gregory, P.J. and B. Marshall, 2012: Attribution of climate change: a methodology to estimate the potential contribution to increases in potato yield in Scotland since 1960. Global Change Biology, 18(4), 1372-1388. Gret-Regamy, A., S. Brunner, J. Altwegg, and P. Bebi, 2013: Facing uncertainties in ecosystem services-based resource management. Journal of Environmental Management, 127(Supplement), S145-S154. Gret-Regamy, A., P. Bebi, I. Bishop, and W. Schmid, 2008: Linking GIS-based models to value ecosystem services in an alpine region. Journal of Environmental Management, 89, 197-208. Grime, J.P., J.D. Fridley, A.P. Askew, K. Thompson, J.G. Hodgson, and C.R. Bennett, 2008: Long-term resistance to simulated climate change in an infertile grassland. Proceedings of the National Academy of Sciences of the United States of America, 105(29), 10028-10032. Grossi, C.M., P. Brimblecombe, and I. Harris, 2007: Predicting long term freeze-thaw risks on Europe built heritage and archaeological sites in a changing climate. Science of the Total Environment, 377(2-3), 273-281. Grossi, C.M., P. Brimblecombe, and H. Lloyd, 2010: The effects of weather on visits to historic properties. Views, 47, 69-71. Grossi, C.M., P. Brimblecombe, B. Mendez, D. Benavente, I. Harris, and M. Deque, 2011: Climatology of salt transitions and implications for stone weathering. Science of the Total Environment, 409(13), 2577-2585. Grossi, M.C., A. Bonazza, P. Brimblecombe, I. Harris, and C. Sabbioni, 2008: Predicting 21st century recession of architectural limestone in European cities. Environmental Geology, 56(3-4), 455-461. Guardiola-Albert, C. and C.R. Jackson, 2011: Potential Impacts of climate change on groundwater supplies to the Donana wetland, Spain. Wetlands, 31(5), 907-920. Guis, H., C. Caminade, C. Calvete, A.P. Morse, A. Tran, and M. Baylis, 2012: Modelling the effects of past and future climate on the risk of bluetongue emergence in Europe. J R Soc Interface., 9(67), 339-350. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Haasnoot, M., H. Middelkoop, A. Offermans, E. van Beek, and W.P.A. van Deursen, 2012: Exploring pathways for sustainable water management in river deltas in a changing environment. Climatic Change, 115(3-4), 795-819. Haddeland, I., P.C. Rhr, and H. Udnaes, 2011: Effects of Climate Changes on Water Resources in the Glomma River Basin - Norway, Technical Report No. 27. Water and Global Change (WATCH), Norway, pp. 1-17. Haigh, I., R. Nicholls, and N. Wells, 2010: Assessing changes in extreme sea levels: Application to the English Channel, 1900-2006. Continental Shelf Research, 30(9), 1042-1055. Haines, A., P. Wilkinson, C. Tonne, and I. Roberts, 2009a: Aligning climate change and public health policies. The Lancet, 374(9707), 2035-2038. Haines, A., A.J. McMichael, K.R. Smith, I. Roberts, J. Woodcock, A. Markandya, B.G. Armstrong, D. Campbell- Lendrum, A.D. Dangour, M. Davies, N. Bruce, C. Tonne, M. Barrett, and P. Wilkinson, 2009b: Public health benefits of strategies to reduce greenhouse-gas emissions: overview and implications for policy makers. The Lancet, 374(9707), 2104-2114. Haines-Young, R., M. Potschin, and F. Kienast, 2012: Indicators of ecosystem service potential at European scales: Mapping marginal changes and trade-offs. Ecological Indicators, 21, 39-53. Hajat, S., M. O'Connor, and T. Kosatsky, 2010: Health effects of hot weather: from awareness of risk factors to effective health protection. Lancet, 375(9717), 856-863. Hakala, K., A.O. Hannukkala, E. Huusela-Veistola, M. Jalli, and P. Peltonen-Sainio, 2011: Pests and diseases in a changing climate: a major challenge for Finnish crop production. Agricultural and Food Science, 20(1), 3-14. Hallegatte, S., F. Henriet, and J. Corfee-Morlot, 2008: The Economics of Climate Change Impacts and Policy Benefits at City Scale: A Conceptual Framework. OECD Environment Working Papers No. 4. OECD Publishing, France, pp. 1-48. Hallegatte, S., N. Ranger, O. Mestre, P. Dumas, J. Corfee-Morlot, C. Herweijer, and R. Wood, 2011: Assessing climate change impacts, sea level rise and storm surge risk in port cities: a case study on Copenhagen. Climatic Change, 104, 113-137. Hallegatte S, Green C, Nicholls RJ, Corfree-Morlot J (2013) Future flood losses in major coastal cities. Nature Climate Change doi:10.1038/nclimate1979. Halpern, B.S., S. Walbridge, K.A. Selkoe, C.V. Kappel, F. Micheli, C. D'Agrosa, J.F. Bruno, K.S. Casey, C. Ebert, H.E. Fox, R. Fujita, D. Heinemann, H.S. Lenihan, E.M.P. Madin, M.T. Perry, E.R. Selig, M. Spalding, R. Steneck, and R. Watson, 2008: A global map of human impact on marine ecosystems. Science, 319, 948-952. Hames, J. and S. Vardoulakis, 2012: Climate Change Risk Assessment for the Health Sector. In: Climate Change Risk Assessment. Department for Environment, Food and Rural Affairs, London, United Kingdom. Hamilton, J.M. and R.S.J. Tol, 2007: The impact of climate change on tourism in Germany, the UK and Ireland: a simulation study. Regional Environmental Change, 7(3), 161-172. Hamin, E.M. and N. Gurran, 2009: Urban form and climate change: Balancing adaptation and mitigation in the U.S. and Australia. Habitat International, 33(3), 238-245. Hamududu, B. and A. Killingtveit, 2012: Assessing climate change impacts on global hydropower. Energies, 5, 305- 322. Hanewinkel, M., D.A. Cullmann, M. Schelhaas, G. Nabuurs, and N.E. Zimmermann, 2013: Climate change may cause severe loss in the economic value of European forest land. Nature Climate Change, 3, 203-207. Hansen, J., R. Ruedy, M. Sato, and K. Lo, 2010: Global surface temperature change. Review of Geophysics, 48(4), RG4004. Hanso, M. and R. Drenkhan, 2007: Retrospective analysis of Lophodermium seditiosum epidemics in Estonia. Acta Silvatica & Lignaria Hungarica, Special Issue, 31-45. Hanson, S., R. Nicholls, N. Ranger, S. Hallegatte, J. Corfee-Morlot, C. Herweijer, and J. Chateau, 2011: A global ranking of port cities with high exposure to climate extremes. Climatic Change, 104(1), 89-111. Hardacre, C.J., P.I. Palmer, K. Baumanns, M. Rounsevell, and D. Murray-Rust, 2012: Probabilistic estimation of future emissions of isoprene and surface oxidant chemistry associated with land use change in response to growing food needs. Atmospheric Chemistry and Physics, Discussion paper 12, 33359-33410. Harrison, G.P., L.C. Cradden, and J.P. Chick, 2008: Preliminary assessment of climate change impacts on the UK onshore wind energy resource. Energy Sources, Part A: Recovery, Utilization, and Environmental Effects, 30(14), 1286-1299. Harrison, P.A., I.P. Holman, G. Cojocaru, K. Kok, A. Kontogianni, M. Metzger, and M. Gramberger, 2013: Combining qualitative and quantitative understanding for exploring cross-sectoral climate change impacts, adaptation and vulnerability in Europe. Regional Environmental Change, 13(4), 761-780. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Harrison, P.A., M. Vandewalle, M.T. Sykes, P.M. Berry, R. Bugter, F. de Bello, C.K. Feld, U. Grandin, R. Harrington, J.R. Haslett, R.H.G. Jongman, G.W. Luck, da Silva P.M., M. Moora, J. Settele, J.P. Sousa, and M. Zobel, 2010: Identifying and prioritising services in European terrestrial and freshwater ecosystems. Biodiversity and Conservation, 19(10), 2791-2821. Harrison, P.A., P.M. Berry, C. Henriques, and I.P. Holman, 2008: Impacts of socio-economic and climate change scenarios on wetlands: Linking water resource and biodiversity meta-models. Climatic Change, 90(1-2), 113- 139. Hartel, T., R. B ncil , and D. Cog lniceanu, 2011: Spatial and temporal variability of aquatic habitat use by amphibians in a hydrologically modified landscape. Freshwater Biology, 56(11), 2288-2298. Hartikainen, K., J. Riikonen, A. Nerg, M. Kivimäenpää, V. Ahonen, A. Tervahauta, S. Kärenlampi, M. Mäenpää, M. Rousi, S. Kontunen-Soppela, E. Oksanen, and T. Holopainen, 2012: Impact of elevated temperature and ozone on the emission of volatile organic compounds and gas exchange of silver birch (Betula pendula Roth). Environmental and Experimental Botany, 84, 33-43. Haugen, J.E. and T. Iversen, 2008: Response in extremes of daily precipitation and wind from a downscaled multi- model ensemble of anthropogenic global climate change scenarios. Tellus Series A Dynamic Meteorology and Oceanography, 60(3), 411-426. Hawkins, E., J. Robson, R. Sutton, D. Smith, and N. Keenlyside, 2011: Evaluating the potential for statistical decadal predictions of sea surface temperatures with a perfect model approach. Climate Dynamics, 37(11-12), 2495-2509. Hawkins, E., T.E. Fricker, A.J. Challinor, C.A.T. Ferro, C.K. Ho, and T.M. Osborne, 2013: Increasing influence of heat stress on French maize yields from the 1960s to the 2030s. Global Change Biology, 19(3), 937-947. Haylock, M.R., N. Hofstra, A.M.G. Klein Tank, E.J. Klok, P.D. Jones, and M. New, 2008: A European daily high- resolution gridded data set of surface temperature and precipitation for 1950-2006. Journal of Geophysical Research, 113, 1-12. Heath, M.R., F.C. Neat, J.K. Pinnegar, D.G. Reid, D.W. Sims, and P.J. Wright, 2012: Review of climate change impacts on marine fish and shellfish around the UK and Ireland. Aquatic Conservation: Marine and Freshwater Ecosystems, 22(3), 337-367. Heidrich, O., R.J. Dawson, D. Reckian, and C.L. Walsh, 2013: Assessment of the climate preparedness of 30 urban areas in the UK Climatic Change, 120(4), 771-784. Hein, L., M.J. Metzger, and A. Moreno, 2009: Potential impacts of climate change on tourism; a case study for Spain. Current Opinion in Environmental Sustainability, 1(2), 170-178. Hekkenberg, M., R. Benders, H. Moll, and A. Schoot Uiterkamp, 2009: Indications for a changing electricity demand pattern: the temperature dependence of electricity in the Netherlands. Energy Policy, 37, 1542-1551. HELCOM, 2007: Climate Change in the Baltic Sea Area HELCOM Thematic Assessment in 2007. In: Baltic Sea Environment Proceedings. Helsinki Commission, Helsinki, Finland, pp. 1-54. Hellmann, F. and J.E. Vermaat, 2012: Impact of climate change on water management in Dutch peat polders. Ecological Modelling, 240, 74-83. Helming, K., K. Diehl, T. Kuhlman, T. Jansson, P.H. Verburg, M. Bakker, M. Perez-Soba, L. Jones, P.J. Verkerk, P. Tabbus, J. Breton Morris, Z. Drillet, J. Farrington, P. LeMouël, P. Zagame, T. Stuczynski, G. Siebielec, S. Sieber, and H. Wiggering, 2011: Ex Ante Impact Assessment of Policies Affecting Land Use, Part B: Application of the Analytical Framework. Ecology and Society, 16(1). Heltberg, R., H. Gitay, and R.G. Prabhu, 2012: Community Based Adaptation: Lessons from a Grant Competition. Climate Policy, 12(2), 143-163. Hemery, G.E., 2008: Forest management and silvicultural responses to projected climate change impacts on European broadleaved trees and forests . International Forestry Review, 10(4), 591-607. Hemery, G.E., J.R. Clark, E. Aldinger, H. Claessens, M.E. Malvolti, E. O'Connor, Y. Raftoyannis, P.S. Savill, and R. Brus, 2010: Growing scattered broadleaved tree species in Europe in a changing climate: a review of risks and opportunities. Forestry, 83(1), 65-81. Henderson, G.R. and D.J. Leathers, 2010: European snow cover extent variability and associations with atmospheric forcings. International Journal of Climatology, 30(10), 1440-1451. Henderson, P.A., 2007: Discrete and continuous change in the fish community of the Bristol Channel in response to climate change. Journal of the Marine Biological Association of the UK, 87(02), 589-589. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hendrickx, F. and E. Sauquet, 2013: Impact of warming climate on water management for the Ariege River basin (France). Hydrological Sciences Journal, 58(5), 976-993 (DOI: 10.1080/02626667.2013.788790). Henriques, C., I.P. Holman, E. Audsley, and K. Pearn, 2008: An interactive multi-scale integrated assessment of future regional water availability for agricultural irrigation in East Anglia and North West England. Climatic Change, 90(1-2), 89-111. Hermans, C.M.L., I.R. Geijzendorffera, F. Ewertb, M.J. Metzgera, P.H. Vereijkene, G.B. Woltjerf, and A. Verhagene, 2010: Exploring the future of European crop production in a liberalised market, with specific consideration of climate change and the regional competitiveness. Ecological Modelling, 221, 2177-2187. Hermant, M., J. Lobry, S. Bonhommeau, J. Poulard, and O. Le Pape, 2010: Impact of warming on abundance and occurrence of flatfish populations in the Bay of Biscay (France). Journal of Sea Research, 64(1-2), 45-53. Hertel, S., A. Le Tertre, K. Jöckel, and B. Hoffmann, 2009: Quantification of the heat wave effect on cause-specific mortality in Essen, Germany. European Journal of Epidemiology, 24(8), 407-414. Herweijer, C., N. Ranger, and R.E.T. Ward, 2009: Adaptation to climate change: Threats and opportunities for the insurance industry. Geneva Papers on Risk and Insurance: Issues and Practice, 34, 360-380. Hilpert, K., F. Mannke, and P. Schmidt-Thome, 2007: Towards Climate Change Adapatation Strategies in the Baltic Sea Region . In: Developing Policies and Adaptation Strategies to Climate Change in the Baltic Sea Region. Geological Survey of Finland, Espoo, Finland. Hinkel, J., R. Nicholls, A. Vafeidis, R. Tol, and T. Avagianou, 2010: Assessing risk of and adaptation to sea-level rise in the European Union: an application of DIVA. Mitigation and Adaptation Strategies for Global Change, 15(7), 703-719. Hlásny, T., Z. Barcza, M. Fabrika, B. Balázs, G. Churkina, J. Pajtík, R. Sedmák, and M. Turèáni, 2011: Climate change impacts on growth and carbon balance of forests in Central Europe. Climate Research, 47(3), 219-236. Hochrainer, S., J. Linnerooth-Bayer, and R. Mechler, 2010: The European Union Solidarity Fund. Mitigation and Adaptation Strategies for Global Change, 15(7), 797-810. Hodzic, A., S. Madronich, B. Bohn, S. Massie, L. Menut, and C. Wiedinmyer, 2007: Wild re particulate matter in Europe during summer 2003: meso-scale modeling of smoke emissions, transport and radiative effects. Atmospheric Chemistry and Physics, 7, 4043-4064. Hoes, O., 2006: Aanpak Wateroverlast in Polders op Basis van Risicobeheer . Technische Universiteit Delft, Delft, Netherlands, 1-188 pp. Hoffmann, B., S. Häfele, and U. Karl, 2013: Analysis of performance losses of thermal power plants in Germany A System Dynamics Model approach using data from regional climate modeling. Energy, 49, 193-203. Hoffmann, I., 2010: Climate change and the characterization, breeding and conservation of animal genetic resources. Animal Genetics, 41(suppl 1), 32-46. Hoinka, K.P., A. Carvalho, and A.I. Miranda, 2009: Regional-scale weather patterns and wildland fires in central Portugal. International Journal of Wildland Fire, 18(1), 36-49. Holland, T. and B. Smit, 2010: Climate change and the wine industry: current research themes and new directions. Journal of Wine Research, 21(2-3), 125-136. House, J.I., H.G. Orr, J.M. Clark, A. Gallego-Sala, C. Freeman, I.C. Prentice, and P. Smith, 2011: Climate change and the British Uplands: evidence for decision-making. Climate Research, 45, 3-12. Howden, N.J.K., T.P. Burt, F. Worrall, M.J. Whelan, and M.Z. Bieroza, 2010: Nitrate concentrations and fluxes in the River Thames over 140 years (1868 - 2008): are increases irreversible? Hydrological Processes, 24, 2657- 2662. Howden, S.M., J.F. Soussana, F.N. Tubiello, N. Chhetri, M. Dunlop, and H. Meinke, 2007: Adapting agriculture to climate change. Proceedings of the National Academy of Sciences of the United States of America, 104(50), 19691-6. HPA, 2012: Health Effects of Climate Change in the UK 2012 Current evidence, recommendations and research gaps. [Vardoulakis, S. and Heaviside, C. (eds.)]. Health Protection Agency, Didcot, United Kingdom, pp. 1- 242. HSY, 2010: Helsinki Metropolitan Area Adaptation to Climate Change Strategy Helsinki Region Environmental Services Authority, Helsinki, Finland, pp. 88. Huang, C., A.G. Barnett, X. Wang, P. Vaneckova, G. FitzGerald, and S. Tong, 2011: Projecting future heat related mortality under climate change scenarios: a systematic review. Environmental Health Perspectives, 119(12), 1681-1990. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hueging, H., R. Haas, K. Born, D. Jacob, and J.G. Pinto, 2013: Regional changes in wind energy potential over Europe using regional climate model ensemble projections. Journal of Applied Meteorology and Climatology, 52(4), 903-917 Huggel, C., N. Salzmann, S. Allen, J. Caplan-Auerbach, L. Fischer, W. Haeberli, C. Larsen, D. Schneider, and R. Wessels, 2010: Recent and future warm extreme events and high-mountain slope stability. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 368(1919), 2435-2459. Huggel, C., J.J. Clague, and O. Korup, 2012: Is climate change responsible for changing landslide activity in high mountains? Earth Surface Processes and Landforms, 37(1), 77-91. Hunt, A. and P. Watkiss, 2011: Climate change impacts and adaptation in cities: A review of the literature. Climatic Change, 104(1), 13-49. Huntjens, P., C. Pahl-Wostl, and J. Grin, 2010: Climate change adaptation in European river basins. Regional Environmental Change, 10, 263-284. Huntley, B., R.E. Green, Y.C. Collingham, and S.G. Willis, 2007: A Climatic Atlas of European Breeding Birds. Lynx Edicions, United Kingdom, pp. 1-834. Hurkmans, R., W. Terink, R. Uijlenhoet, P. Torfs, D. Jacob, and P.A. Troch, 2010: Changes in streamflow dynamics in the Rhine basin under three high-resolution regional climate scenarios. Journal of Climate, 23(3), 679-699. Huss, M., 2011: Present and future contribution of glacier storage change to runoff from macroscale drainage basins in Europe. Water Resources Research, 47(7), (doi:10.1029/2010WR010299). ICES, 2010: Book 6: North Sea. Cod in Subarea IV. In: Report of the ICES Advisory Committee 2010. ICES Advice, 2010. International Council for the Exploration of the Sea, Copenhagen, Denmark, pp. 1-309. Iglesias, A., L. Garrote, F. Flores, and M. Moneo, 2007: Challenges to Manage the Risk of Water Scarcity and Climate Change in the Mediterranean. Water Resources Management, 21(5), 775-788. Iglesias, A., L. Garrote, A. Diz, J. Schlickenrieder, and M. Moneo, 2012: Water and people: Assessing policy priorities for climate change adaptation in the Mediterranean. In: Regional Assessment of Climate Change in the Mediterranean. [Navarra, A. and L. Tubiana (eds.)]. Springer, Dordrecht, Netherlands, pp. 201-233. Isaac, M. and D.P. van Vuuren, 2009: Modelling global residential sector energy demand for heating and air conditioning in the context of climate change. Energy Policy, 37(2), 507-521. Jackson, A.C. and J. McIlvenny, 2011: Coastal squeeze on rocky shores in northern Scotland and some possible ecological impacts. Journal of Experimental Marine Biology and Ecology, 400(1-2), 314-321. Jackson, C.R., R. Meister, and C. Prudhomme, 2011: Modelling the effects of climate change and its uncertainty on UK Chalk groundwater resources from an ensemble of global climate model projections. Journal of Hydrology, 399(1-2), 12-28. Jacob, D. and R. Podzun, 2010: Global warming below 2°C relative to pre-industrial level: how might climate look like in Europe. Nova Acta Leopoldina, 384, 71-76. Jacob, D., J. Petersen, B. Eggert, A. Alias, O. Bossing Christensen, L.M. Bouwer, A. Braun, A. Colette, M. Deque, G. Georgievski, E. Georgopoulou, A. Gobiet, G. Nikulin, A. Haensler, N. Hempelmann, C. Jones, K. Keuler, S. Kovats, N. Kröner, S. Kotlarski, A. Kriegsmann, E. Martin, E. van Meijgaard, C. Moseley, S. Pfeifer, S. Preuschmann, K. Radtke, D. Rechid, M. Rounsevell, P. Samuelsson, S. Somot, J. Soussana, C. Teichmann, R. Valentini, R. Vautard, and B. Weber, 2013: EURO-CORDEX: New high-resolution climate change projections for European impact research. Regional Environmental Change, 1-16 (in press). Jacob, D.J. and D.A. Winner, 2009: Effect of climate change on air quality. Atmospheric Environment, 43(1), 51-63. Jactel, H., B.C. Nicoll, M. Branco, J. Gonzalez-Olabarria, W. Grodzki, B. Langström, F. Moreira, S. Netherer, C. Orazio, D. Piou, H. Santos, M.J. Schelhaas, K. Tojic, and F. Vodde, 2009: The influences of forest stand management on biotic and abiotic risks of damage. Annals of Forest Science, 66(7), 1-18. Jacxsens, L., P.A. Luning, J.G.A.J. van der Vorst, F. Devlieghere, R. Leemans, and M. Uyttendaele, 2010: Simulation modelling and risk assessment as tools to identify the impact of climate change on microbiological food safety-The case study of fresh produce supply chain. Food Research International, 43(7), 1925-1935. James, P., K. Tzoulas, M.D. Adams, A. Barber, J. Box, J. Breuste, T. Elmqvist, M. Frith, C. Gordon, K.L. Greening, J. Handley, S. Haworth, A.E. Kazmierczak, M. Johnston, K. Korpela, M. Moretti, J. Niemelä, S. Pauleit, M.H. Roe, J.P. Sadler, and C. Ward Thompson, 2009: Towards an integrated understanding of green space in the European built environment. Urban Forestry and Urban Greening, 8(2), 65-75. Jenkins, D., Y. Liu, and A.D. Peacock, 2008: Climatic and internal factors affecting future UK office heating and cooling energy consumptions. Energy and Buildings, 40(5), 874-881. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Jenkins, D.P., 2009: The importance of office internal heat gains in reducing cooling loads in a changing climate. International Journal of Low-Carbon Technologies, 4(3), 134-140. Jeppesen, E., B. Kronvang, J.E. Olesen, J. Audet, M. Sndergaard, C.C. Hoffmann, H.E. Andersen, T.L. Lauridsen, L. Liboriussen, S.E. Larsen, M. Beklioglu, M. Meerhoff, A. Özen, and K. Özkan, 2011: Climate change effects on nitrogen loading from cultivated catchments in Europe: implications for nitrogen retention, ecological state of lakes and adaptation. Hydrobiologia, 663(1), 1-21. Jiguet, F., R.D. Gregory, V. Devictor, R.E. Green, P. Vorisek, A. Van Strien, and D. Couvet, 2010: Population trends of European common birds are predicted by characteristics of their climatic niche. Global Change Biology, 16(2), 497-505. Johnk, K.D., J. Huisman, J. Sharples, B. Sommeijer, P.M. Visser, and J.M. Stroom, 2008: Summer heatwaves promote blooms of harmful cyanobacteria. Global Change Biology, 14(3), 495-512. Johnson, A., M. Acreman, M. Dunbar, S. Feist, A. Giacomello, R. Gozlan, S. Hinsley, A. Ibbotson, H. Jarvie, J. Jones, M. Longshaw, S. Maberly, T. Marsh, C. Neal, J. Newman, M. Nunn, R. Pickup, N. Reynard, C. Sullivan, J. Sumpter, and R. Williams, 2009: The British river of the future: How climate change and human activity might affect two contrasting river ecosystems in England. Science of the Total Environment, 407, 4787-4798. Jomelli, V., D. Brunstein, D. Grancher, and P. Pech, 2007: Is the response of hill slope debris flows to recent climate change univocal? A case study in the Massif des Ecrins (French Alps). Climatic Change, 85(1-2), 119-137. Jomelli, V., D. Brunstein, M. Déqué, M. Vrac, and D. Grancher, 2009: Impacts of future climatic change (2070- 2099) on the potential occurrence of debris flows: A case study in the Massif des Ecrins (French Alps) . Climatic Change, 97(1), 171-191. Jonkeren, O., P. Rietveld, and J. van Ommeren, 2007: Climate change and inland waterway transport; welfare effects of low water levels on the river Rhine. Journal of Transport Economics and Policy, 41(3), 387-411. Jonkeren, O., B. Jourquin, and P. Rietveld, 2011: Modal-split effects of climate change: The effect of low water levels on the competitive position of inland waterway transport in the River Rhine area. Transportation Research Part A: Policy & Practice, 45(10), 1007-1019. Jonkeren, O.E., 2009: Adaptation to Climate Change in Inland Waterway Transport. Diss. Ph.D., Vrije Universiteit, Amsterdam, Netherlands, 1-164 pp. Jönsson, A.M., S. Harding, P. Krokene, H. Lange, A. Lindelöw, B. Okland, H.P. Ravn, and L.M. Schroeder, 2011: Modelling the potential impact of global warming on Ips typographus voltinism and reproductive diapause . Climatic Change, 109(3-4), 695-718. Jönsson, A.M., G. Appelberg, S. Harding, and L. Bärring, 2009: Spatio-temporal impact of climate change on the activity and voltinism of the spruce bark beetle, Ips typographus. Global Change Biology, 15(2), 486-499. Jonsson, B. and N. Jonsson, 2009: A review of the likely effects of climate change on anadromous Atlantic salmon Salmo salar and brown trout Salmo trutta, with particular reference to water temperature and flow. Journal of Fish Biology, 75(10), 2381-2447. Jonzén, N., A. Lindén, T. Ergon, E. Knudsen, J.O. Vik, D. Rubolini, D. Piacentini, C. Brinch, F. Spina, L. Karlsson, M. Stervander, A. Andersson, J. Waldenström, A. Lehikoinen, E. Edvardsen, R. Solvang, and N.C. Stenseth, 2006: Rapid advance of spring arrival dates in long-distance migratory birds. Science, 312(5782), 1959-1961. Jorda, G., D. Gomis, and M. Marcos, 2012: Comment on "Storm surge frequency reduction in Venice under climate change" by Troccoli et al. Climatic Change, 113(3-4), 1081-1087. JRC, 2008: Forest Fires in Europe 2007. Technical Report No.8. EUR 23492 EN - 2008. Joint Research Centre, Institute for Environment and Sustainability, Luxembourg. JRC-EEA, 2010: The European Environment, State and Outlook 2010. Soil. European Environment Agency, Copenhagen, Denmark, pp. 1-44. Kabat, P., L.O. Fresco, M.J.F. Stive, C.P. Veerman, J.S.L.J. van Alphen, B.W.A.H. Parmet, W. Hazeleger, and C.A. Katsman, 2009: Dutch coasts in transition. Nature Geosciences, 2, 450-452. Karaca, M. and R.J. Nicholls, 2008: Potential Implications of Accelerated Sea-Level Rise for Turkey. Journal of Coastal Research, 24(2), 288-298. Katsman, C., A. Sterl, J. Beersma, H. van den Brink, J. Church, W. Hazeleger, R. Kopp, D. Kroon, J. Kwadijk, R. Lammersen, J. Lowe, M. Oppenheimer, H. Plag, J. Ridley, H. von Storch, D. Vaughan, P. Vellinga, L. Vermeersen, R. van de Wal, and R. Weisse, 2011: Exploring high-end scenarios for local sea level rise to develop flood protection strategies for a low-lying delta: the Netherlands as an example. Climatic Change, 109, 617-645. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kay, A.L., S.M. Crooks, P. Pall, and D.A. Stone, 2011: Attribution of Autumn/Winter 2000 flood risk in England to anthropogenic climate change: A catchment-based study. Journal of Hydrology, 406(1 2), 97-112. Keenan, T., J. Maria Serra, F. Lloret, M. Ninyerola, and S. Sabate, 2011: Predicting the future of forests in the Mediterranean under climate change, with niche- and process-based models: CO2 matters! Global Change Biology, 17(1), 565-579. Keith, S.A., A.C. Newton, R.J.H. Herbert, M.D. Morecroft, and C.E. Bealey, 2009: Non-analogous community formation in response to climate change. Journal for Nature Conservation, 17(4), 228-235. Kersebaum, K.C., A.S. Nain, C. Nendel, M. Gandorfer, and M. Wegehenkel, 2008: Simulated effect of climate change on wheat production and nitrogen management at different sites in Germany. Journal of Agrometeorology, 10, 266-273. Keskitalo, E.C.H., 2010: The Development of Adaptation Policy and Practice in Europe: Multi-level Governance of Climate Change. Springer, Dordrecht., pp. 376. Khaledian, M.R., J.C. Mailhol, P. Ruelle, I. Mubarak, and S. Perret, 2010: The impacts of direct seeding into mulch on the energy balance of crop production system in the SE of France. Soil and Tillage Research, 106(2), 218- 226. Kilpeläinen, M. and H. Summala, 2007: Effects of weather and weather forecasts on driver behaviour. Transportation Research, 10(4), 288-299. Kint, V., W. Aertsen, M. Campioli, D. Vansteenkiste, A. Delcloo, and B. Muys, 2012: Radial growth change of temperate tree species in response to altered regional climate and air quality in the period 1901-2008. Climatic Change, 115, 343-363. Kjellström, E., G. Nikulin, U. Hansson, G. Strandberg, and A. Ullerstig, 2011: 21st century changes in the European climate: uncertainties derived from an ensemble of regional climate model simulations. Tellus, 63A(1), 24-40. Kjellstrom, T., R.S. Kovats, S.J. Lloyd, T. Holt, and R.S. Tol, 2009: The direct impact of climate change on regional labor productivity. Archives of Environmental and Occupational Health, 64(4), 217-227. Klaus, M., A. Holsten, P. Hostert, and J.P. Kropp, 2011: Integrated methodology to assess windthrow impacts on forest stands under climate change. Forest Ecology and Management, 261(11), 1799-1810. Klijn, F., N. Asselman, and H. Van Der Most, 2009: Compartmentalisation: flood consequence reduction by splitting up large polder areas. Journal of Flood Risk Management, 3, 3-17. Klik, A. and J. Eitzinger, 2010: Impact of climate change on soil erosion and the efficiency of soil conservation practices in Austria. Journal of Agricultural Science, 148, 529-541. Klve, B., P. Ala-aho, G. Bertrand, Z. Boukalova, A. Ertürk, N. Goldscheider, J. Ilmonen, N. Karakaya, H. Kupfersberger, J. Kv rner, A. Lundberg, M. Mileusniæ, A. Moszczynska, T. Muotka, E. Preda, P. Rossi, D. Siergieiev, J. ©imek, P. Wachniew, V. Angheluta, and A. Widerlund, 2011: Groundwater dependent ecosystems. Part I: Hydroecological status and trends. Environmental Science and Policy, 14(7), 770-781. Koca, D., B. Smith, and M. Sykes, 2006: Modelling regional climate change effects on potential natural ecosystems in Sweden. Climatic Change, 78, 381-406. Koch, H. and S. Vögele, 2009: Dynamic modeling of water demand, water availability and adaptation strategies for power plants to global change. Ecological Economics, 68(7), 2031-2039. Koetse, M.J. and P. Rietveld, 2009: The impact of climate change and weather on transport: An overview of empirical findings. Transportation Research, 14(3), 205-221. Koutroulis, A.G., A. Vrochidou, and I.K. Tsanis, 2010: Spatiotemporal Characteristics of Meteorological Drought for the Island of Crete. Journal of Hydrometeorology, 12(2), 206-226 (doi: 10.1175/2010JHM1252.1). Koutsias, N., M. Arianoutsou, A.S. Kallimanis, G. Mallinis, J.M. Halley, and P. Dimopoulos, 2012: Where did the fires burn in Peloponnisos, Greece the summer of 2007? Evidence for a synergy of fuel and weather. Agricultural and Forest Meteorology, 156, 41-53. Kovats, R.S. and S. Hajat, 2008: Heat Stress and Public Health: A Critical Review. Annual Review of Public Health, 29, 41-55. Kreienkamp, F., A. Spekat, and W. Enke, 2010: Stationarity of atmospheric waves and blocking over Europe based on a reanalysis dataset and two climate scenarios. Theoretical and Applied Climatology, 102(1-2), 205- 212 (DOI 10.1007/s00704-010-0261-3). Krekt, A.H., T.J. van der Laan, R.A.E. van der Meer, B. Turpijn, O.E. Jonkeren, A. van der Toorn, E. Mosselman, J. van Meijeren, and T. Groen, 2011: Climate Change and Inland Waterway Transport: Impacts on the Sector, the Port of Rotterdam and Potential Solutions. Kennis voor Klimaat, Netherlands, pp. 1-76. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kriegler, E., B.C. O'Neill, S. Hallegatte, T. Kram, R.H. Moss, R. Lempert, and T.J. Wilbanks, 2010: Socio- Economic Scenario Development for Climate Change Analysis. In: CIRED Working Paper, DT/WPNo 2010 23. CIRED, France, pp. 1-35. Kristensen, K., K. Schelde, and J.E. Olesen, 2011: Winter wheat yield response to climate variability in Denmark. Journal of Agricultural Science, 149(1), 33-47. Kuhlicke, C., A. Steinführer, C. Begg, C. Bianchizza, M. Bründl, M. Buchecker, B. De Marchi, M. Di Masso Tarditti, C. Höppner, B. Komac, L. Lemkow, J. Luther, S. Mccarthy, L. Pellizzoni, O. Renn, A. Scolobig, M. Supramaniam, S. Tapsell, G. Wachinger, G. Walker, R. Whittle, M. Zorn, and H. Faulkner, 2011: Perspectives on social capacity building for natural hazards: Outlining an emerging field of research and practice in Europe. Environmental Science and Policy, 14(7), 804-814. Kundzewicz, Z.W., I. Piñskwar, and G.R. Brakenridge, 2013: Large floods in Europe, 1985-2009. Hydrological Sciences Journal, 58(1), 1-7. Kunz, M., J. Sander, and C. Kottmeier, 2009: Recent trends of thunderstorm and hailstorm frequency and their relation to atmospheric characteristics in southwest Germany. International Journal of Climatology, 29, 2283- 2297. Kwadijk, J.C.J., M. Haasnoot, J.P.M. Mulder, M.M.C. Hoogvliet, A.B.M. Jeuken, R.A.A. van der Krogt, N.G.C. van Oostrom, H.A. Schelfhout, E.H. van Velzen, H. van Waveren, and M.J.M. de Wit, 2010: Using adaptation tipping points to prepare for climate change and sea level rise, a case study in the Netherlands. Wiley Interdisciplinary Reviews: Climate Change, 1(5), 729-740. Ladanyi, M., 2008: Risk methods and their applications in agriculture. Applied Ecology and Environmental Research, 6(1), 147-164. Lake, I.R., I.A. Gillespie, G. Bentham, G.L. Nichols, C. Lane, G.K. Adak, and E.J. Threlfall, 2009: A re-evaluation of the impact of temperature and climate change on foodborne illness. Epidemiology and Infection, 137(11), 1538-1547. Lal, R., J.A. Delgado, P.M. Groffman, N. Millar, C. Dell, and A. Rotz, 2011: Management to mitigate and adapt to climate change. Journal of Soil and Water Conservation, 66(4), 276-285. Lamond, J.E., D.G. Proverbs, and F.N. Hammond, 2009: Accessibility of flood risk insurance in the UK: confusion, competition and complacency. Journal of Risk Research, 12(6), 825-841. Lankester, P. and P. Brimblecombe, 2010: Predicting future indoor climate at Knole. Views, 47, 71-73. Lasda, O., A. Dikou, and E. Papapanagiotou, 2010: Flash Flooding in Attika, Greece: Climatic Change or Urbanization? Ambio, 39, 608-611. Lasserre, F. and S. Pelletier, 2011: Polar super seaways? Maritime transport in the Arctic: An analysis of shipowners' intentions. Journal of Transport Geography, 19, 1465-1473. Lavalle, C., F. Micale, T.D. Houston, A. Camia, R. Hiederer, C. Lazar, C. Conte, G. Amatulli, and G. Genovese, 2009: Climate change in Europe. 3. Impact on agriculture and forestry. A review. Agronomy for Sustainable Development, 29(3), 433-446. Lawrence, D. and H. Hisdal, 2011: Hydrological Projections for Floods in Norway Under a Future Climate. NVE Report no. 2011-5. Norwegian Water Resources and Energy Directorate, Norway, pp. 1-54. Le Floc'h, P.L., J. Poulard, O. Thébaud, F. Blanchard, J. Bihel, and F. Steinmetz, 2008: Analyzing the market position of fish species subject to the impact of long-term changes: A case study of French fisheries in the Bay of Biscay. Aquatic Living Resources, 21(3), 307-316. Leander, R., T.A. Buishand, B.J.J.M. van den Hurk, and M.J.M. de Wit, 2008: Estimated changes in flood quantiles of the river Meuse from resampling of regional climate model output. Journal of Hydrology, 351(3-4), 331-343. Lee, H.C., R. Walker, S. Haneklaus, L. Philips, G. Rahmann, and E. Schnug, 2008: Organic farming in Europe: A potential major contribution to food security in a scenario of climate change and fossil fuel depletion. Landbauforschung Volkenrode, 58(3), 145-151. Lejeusne, C., P. Chevaldonne, C. Pergent-Martini, C.F. Boudouresque, and T. Perez, 2009: Climate change effects on a miniature ocean: the highly diverse, highly impacted Mediterranean Sea. Trends in Ecology and Evolution, 25(4). Lemoine, N., H. Schaefer, and K. Böhning-Gaese, 2007a: Species richness of migratory birds is influenced by global climate change. Global Ecology and Biogeography, 16(1), 55-65. Lemoine, N., H. Bauer, M. Peintinger, and K. Böhning-Gaese, 2007b: Effects of climate and land-use change on species abundance in a Central European bird community. Conservation Biology, 21(2), 495-503. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Lemonsu, A., R. Kounkou-Arnaud, J. Desplat, J. Salagnac, and V. Masson, 2013: Evolution of the Parisian urban climate under a global changing climate. Climatic Change, 116(3-4), 679-692. Lenderink, G. and E. Van Meijgaard, 2008: Increase in hourly precipitation extremes beyond expectations from temperature changes. Nature Geoscience, 1(8), 511-514. Lennert, M. and J. Robert, 2010: The territorial futures of Europe: 'Trends', 'Competition' or 'Cohesion'. Futures, 42(8), 833-845. Lenoir, J., J.C. Gegout, P.A. Marquet, P. de Ruffray, and H. Brisse, 2008: A significant upward shift in plant species optimum elevation during the 20th century. Science, 320(5884), 1768-1771. Lenoir, S., G. Beaugrand, and É. Lecuyer, 2011: Modelled spatial distribution of marine fish and projected modifications in the North Atlantic Ocean. Global Change Biology, 17(1), 115-129. Letourneau, A., P.H. Verburg, and E. Stehfest, 2012: A land-use systems approach to represent land-use dynamics at continental and global scales. Environmental Modelling and Software, 33, 61-79. Levinsky, I., F. Skov, J. Svenning, and C. Rahbek, 2007: Potential impacts of climate change on the distributions and diversity patterns of European mammals. Biodiversity and Conservation, 16(13), 3803-3816. Liberloo, M., S. Luyssaert, V. Bellassen, S.N. Djomo, M. Lukac, C. Calfapietra, I.A. Janssens, M.R. Hoosbeek, N. Viovy, G. Churkina, G. Scarascia-Mugnozza, and R. Ceulemans, 2010: Bio-energy retains its mitigation potential under elevated CO2. Public Library on Science, 5(7), e11648 (doi:10.1371/journal.pone.0011648). Linard, C., N. Poncon, D. Fontenille, and E.F. Lambin, 2009: Risk of malaria reemergence in Southern France: testing scenarios with a multiagent simulation model. Eohealth, 6(1), 135-147. Lindgren, J., D.K. Johnsson, and A. Carlsson-Kanyama, 2009: Climate adaptation of railways: Lessons from Sweden. European Journal of Transport and Infrastructure Research, 9(2), 164-181. Lindner, M., M. Maroschek, S. Netherer, A. Kremer, A. Barbati, J. Garcia-Gonzalo, R. Seidl, S. Delzon, P. Corona, M. Kolström, M.J. Lexer, and M. Marchetti, 2010: Climate change impacts, adaptive capacity, and vulnerability of European forest ecosystems. Forest Ecology and Management, 259(4), 698-709. Lindsay, S.W., D.G. Hole, R.A. Hutchinson, S.A. Richards, and S.G. Willis, 2010: Assessing the future threat from vivax malaria in the United Kingdom using two markedly different modelling approaches. Malaria Journal, 9(1), 70-78. Linnerud, K., T.H. Mideska, and G.S. Eskeland, 2011: The impact of climate change on nuclear power supply. The Energy Journal, 32(1), 149-168. Lionello, P., M.B. Galati, and E. Elvini, 2012: Extreme storm surge and wind wave climate scenario simulations at the Venetian littoral. Physics and Chemistry of the Earth, 40-41, 86-92. Liu, M. and J. Kronbak, 2010: The potential economic viability of using the Northern Sea Route (NSR) as an alternative route between Asia and Europe. Journal of Transport Geography, 18, 434-444. Liu, Y., R.A. Kahn, A. Chaloulakou, and P. Koutrakis, 2009: Analysis of the impact of the forest res in August 2007 on air quality of Athens using multi-sensor aerosol remote sensing data, meteorology and surface observations. Atmospheric Environment, 43, 3310-3318. Lobell, D.B., W. Schlenker, and J. Costa-Roberts, 2011: Climate trends and global crop production since 1980. Science, 333(6042), 616-620. Long, S.P., E.A. Ainsworth, A.D.B. Leakey, J. Nosberger, and D.R. Ort, 2006: Food for Thought: Lower-Than- Expected Crop Yield Stimulation with Rising CO2 Concentrations. Science, 312(5782), 1918-1921. Lopez Saez, J., C. Corona, M. Stoffel, and F. Berger, 2013: Climate change increases frequency of shallow spring landslides in the French Alps. Geology, 41(5), 619-622. Lorz, C., C. Furst, Z. Galiz, D. Matijasic, V. Podrazky, N. Potocic, P. Simoncic, M. Strauch, H. Vacik, and F. Makeschin, 2010: GIS-based Probability Assessment of Natural Hazards in Forested Landscapes of Central and South-Eastern Europe. Environmental Management, 46, 920-930. Lowe, J.A., T.P. Howard, A. Pardaens, J. Tinker, J. Holt, S. Wakelin, G. Milne, J. Leake, J. Wolf, K. Horsburgh, T. Reeder, G. Jenkins, J. Ridley, S. Dye, and S. Bradley, 2009: UK Climate Projections Science Report: Marine and Coastal Projections. Met Office Hadley Centre, Exeter, United Kingdom, pp. 99. Lowe, D., K.L. Ebi, and B. Forsberg, 2011: Heatwave early warning systems and adaptation advice to reduce human health consequences of heatwaves. International Journal of Environmental Research and Public Health, 8(12), 4623-4648. Luck, J., M. Spackman, A. Freeman, P. Trebicki, W. Griffiths, K. Finlay, and S. Chakraborty, 2011: Climate change and diseases of food crops. Plant Pathology, 60(1), 113-121. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ludwig, R., R. Roson, C. Zografos, and G. Kallis, 2011: Towards an inter-disciplinary research agenda on climate change, water and security in Southern Europe and neighboring countries. Environmental Science and Policy, 14(7), 794-803. Lugeri, N., Z. Kundzewicz, E. Genovese, S. Hochrainer, and M. Radziejewski, 2010: River flood risk and adaptation in Europeóassessment of the present status. Mitigation and Adaptation Strategies for Global Change, 15(7), 621-639. Lung, T., C. Lavalle, R. Hiederer, A. Dosio, and L.M. Bouwer, 2012: A multi-hazard regional level impact assessment for Europe combining indicators of climatic and non-climatic change. Global Environmental Change, 23(2), 522-536. Luterbacher, J., D. Dietrich, E. Xoplaki, M. Grosjean, and H. Wanner, 2004: European seasonal and annual temperature variability, trends, and extremes since 1500. Science, 303(5663), 1499-1503. Lyons, S., K. Mayor, and R. Tol, 2009: Holiday destinations: Understanding the travel choices of Irish tourists. Tourism Management, 30(5), 683-692. Maaskant, B., S.N. Jonkman, and L.M. Bouwer, 2009: Future risk of flooding: an analysis of changes in potential loss of life in South Holland (the Netherlands). Environmental Science & Policy, 12(2), 157-169. Mackenzie, B.R., H. Gislason, C. Mollmann, and F.W. Koster, 2007: Impact of 21st century climate change on the Baltic Sea fish community and fisheries. Global Change Biology, 13(7), 1348-1367. Macleod, C.J.A., P.D. Falloon, R. Evans, and P.M. Haygarth, 2012: The Effects of Climate Change on the Mobilization of Diffuse Substances from Agricultural Systems. Advances in Agronomy, 115, 41-47. Madgwick, J.W., J.S. West, R.P. White, M.A. Semenov, J.A. Townsend, J.A. Turner, and B.D.L. Fitt, 2011: Impacts of climate change on wheat anthesis and fusarium ear blight in the UK. European Journal of Plant Pathology, 130(1), 117-131. Magnan, A., B. Garnaud, R. Billé, F. Gemenne, and S. Hallegatte, 2009: The Future of the Mediterranean: From Impacts of Climate Change to Adaptation Issues. Institut du développement durable et des relations internationales (IDDRI), Paris, France, pp. 1-43. Majone, B., C.I. Bovolo, A. Bellin, S. Blenkinsop, and H.J. Fowler, 2012: Modeling the impacts of future climate change on water resources for the Gállego river basin (Spain). Water Resources Research, 48(1), W01512. Malheiro, A.C., J.A. Santos, H. Fraga, and J.G. Pinto, 2010: Climate change scenarios applied to viticultural zoning in Europe. Climate Research, 43, 163-177. Mandryk, M., P. Reidsma, and M. Ittersum, 2012: Scenarios of long-term farm structural change for application in climate change impact assessment. Landscape Ecology, 27(4), 509-527. Mantyka-pringle, C.S., T.G. Martin, and J.R. Rhodes, 2012: Interactions between climate and habitat loss effects on biodiversity: A systematic review and meta-analysis. Global Change Biology, 18(4), 1239-1252. Marcais, B. and M. Desprez-Loustau, 2007: Le réchauffement climatique a-t-il un impact sur les maladies forestieres? RenDez-Vous Techniques, 3, 47-52. Marcos-Lopez, M., P. Gale, B.C. Oidtmann, and E.J. Peeler, 2010: Assessing the impact of climate change on disease emergence in freshwater fish in the United Kingdom. Transboundary and Emerging Diseases, 57(5), 293-304. Marker, M., L. Angeli, L. Bottai, R. Costantini, R. Ferrari, L. Innocenti, and G. Siciliano, 2008: Assessment of land degradation susceptibility by scenario analysis: A case study in Southern Tuscany, Italy. Geomorphology, 93(1- 2), 120-129. Marmot, M., J. Allen, R. Bell, E. Bloomer, P. Goldblatt, and Consortium for the European Review of Social Determinants of Health and the Health Divide., 2012: WHO European review of social determinants of health and the health divide. Lancet, 380(9846), 1011-1029. Marques, S., J.G. Borges, J. Garcia-Gonzalo, F. Moreira, J.M.B. Carreiras, M.M. Oliveira, A. Cantarinha, B. Botequim, and J.M.C. Pereira, 2011: Characterization of wildfires in Portugal. European Journal of Forest Research, 130(5), 775-784. Mavrogianni, A., P. Wilkinson, M. Davies, P. Biddulph, and E. Oikonomou, 2012: Building characteristics as determinants of propensity to high indoor summer temperatures in London dwellings. Building and Environment, 55, 117-130. Mazaris, A.D., A.D. Papanikolaou, M. Barbet-Massin, A.S. Kallimanis, F. Jiguet, D.S. Schmeller, and J.D. Pantis, 2013: Evaluating the Connectivity of a Protected Areas' Network under the Prism of Global Change: The Efficiency of the European Natura 2000 Network for Four Birds of Prey. PLoS ONE, 8(3), e59640 (doi:10.1371/journal.pone.0059640). Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McCarthy, M., M. Best, and R. Betts, 2010: Climate change in cities due to global warming and urban effects. Geophysical Research Letters, 37(9) (DOI: 10.1029/2010GL042845). McColl, L., E. Palin, H. Thornton, D. Sexton, R. Betts, and K. Mylne, 2012: Assessing the potential impact of climate change on the UK's electricity network. Climatic Change, 115(3-4), 821-835. McGrath, J. and D.B. Lobell, 2011: An independent method of deriving the carbon dioxide fertilization effect in dry conditions using historical yield data from wet and dry years. . Global Change Biology, 17, 2689-2696. McHugh, M., 2007: Short-term changes in upland soil erosion in England and Wales: 1999 to 2002. Geomorphology, 86(1-2), 204-213. McInnes, K.L., T.A. Erwin, and J.M. Bathols, 2011: Global Climate Model projected changes in 10 m wind speed and direction due to anthropogenic climate change. Atmospheric Science Letters, 12(4), 325-333. Mechler, R., S. Hochrainer, A. Aaheim, H. Salen, and A. Wreford, 2010: Modelling economic impacts and adaptation to extreme events: Insights from European case studies. Mitigation and Adaptation Strategies for Global Change, 15(7), 737-762. Medlock, J.M. and A.G.C. Vaux, 2013: Colonization of UK coastal realignment sites by mosquitoes: implications for design, management, and public health. Journal of Vector Ecology, 38(1), 53-62. Medri, S., E. Banos de Guisasola, and S. Gualdi, 2012: Overview of the main international climate services. CMCC Research Paper. Issue RP0134. . Euro-Mediterranean Center on Climate Change. Medri, S., S. Venturini, and S. Castellari, 2013: Overview of key climate change impacts, vulnerabilities and adaptation action in Italy. In: CMCC Research Papers. RP0178, pp. 75. Melchiorre, C. and P. Frattini, 2012: Modelling probability of rainfall-induced shallow landslides in a changing climate, Otta, Central Norway. Climatic Change, 113(2), 413-436. Meleux, F., F. Solmon, and F. Giorgi, 2007: Increase in summer European ozone amounts due to climate change. Atmospheric Environment, 41(35), 7577-7587. Menendez, M. and P.L. WoodWorth, 2010: Changes in extreme high water levels based on quasi global tide-gauge data set. Journal of Geophysical Research, 115. Menzel, A., T.H. Sparks, N. Estrella, E. Koch, A. Aasa, A. Ahas K., P. Bissolli, O. Braslavská, A. Briede, F.M. Chmielewski, Z. Crepinsek, Y. Curnel, A. Dahl, C. Defila, A. Donnelly, Y. Filella, K. Jatczak, F. Mage, A. Mestre, O. Nordli, J. Penuelas, P. Pirinen, V. Remi¹ová, H. Scheifinger, M. Striz, A. Susnik, A. Van vliet, F. Wielgolaski, S. Zach, and A. Zust, 2006: European phenological response to climate change matches the warming pattern. Global Change Biology, 12(10), 1969-1976. Metzger, M.J. and M.D.A. Rounsevell, 2011: A need for planned adaptation to climate change in the wine industry. Environmental Research Letters, 6(3), 031001 (doi:10.1088/1748-9326/6/3/031001). Metzger, M.J., D. Schroter, R. Leemans, and W. Cramer, 2008: A spatially explicit and quantitative vulnerability assessment of ecosystem service change in Europe. Regional Environmental Change, 8(3), 91-107. Metzger, M.J., R.G.H. Bunce, R.H.G. Jongman, C.A. Mücher, and J.W. Watkins, 2005: A climatic stratification of the environment of Europe. Global Ecology and Biogeography, 14(6), 549-563. Michelozzi, P., G. Accetta, M. De Sario, D. D'Ippoliti, C. Marino, M. Baccini, A. Biggeri, H.R. Anderson, K. Katsouyanni, F. Ballester, L. Bisanti, E. Cadum, B. Forsberg, F. Forastiere, P.G. Goodman, A. Hojs, U. Kirchmayer, S. Medina, A. Paldy, C. Schindler, J. Sunyer, C.A. Perucci, and PHEWE Collaborative Group, 2009: High temperature and hospitalizations for cardiovascular and respiratory causes in 12 European cities. American Journal of Respiratory and Critical Care Medicine, 179(5), 383-389. Mickwitz, P., S. Beck, A. Jensenm, A.B. Pedersen, C. Görg, M. Melanen, N. Ferrand, C. Kuhlicke, W. Kuindersma, M. Mánez, H. Reinert, and S. Bommel, 2009: Climate policy integration as a necessity for an efficient climate policy. IOP Conf. Series: Earth and Environmental Science, 6, 582017 (doi:10.1088/1755-1307/6/58/582017). Mieszkowska, N., M.J. Genner, S.J. Hawkins, and D.W. Sims, 2009: Chapter 3. Effects of climate change and commercial fishing on Atlantic cod Gadus morhua. Advances in Marine Biology, 56, 213-273. Milad, M., H. Schaich, M. Bürgi, and W. Konold, 2011: Climate change and nature conservation in Central European forests: a review of consequences, concepts and challenges. Forest Ecology and Management, 261(4), 829-843. Millar, C.I., N.L. Stephenson, and S.L. Stephens, 2007: Climate change and forests of the future: Managing in the face of uncertainty. Ecological Applications, 17(8), 2145-2151. Milner, J., M. Davies, and P. Wilkinson, 2012: Urban energy, carbon management (low carbon cities) and co- benefits for human health. Current Opinion in Environmental Sustainability, 4(4), 338-404. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ministry of Agriculture and Forestry, 2009: Evaluation of the Implementation of Finland s National Strategy for Adaptation to Climate Change 2009. Ministry of Agriculture and Forestry, Helsinki, Finland, pp. 1-45. Miraglia, M., H.J.P. Marvin, G.A. Kleter, P. Battilani, C. Brera, E. Coni, F. Cubadda, L. Croci, B. De Santis, S. Dekkers, L. Filippi, R.W.A. Hutjes, M.Y. Noordam, M. Pisante, G. Piva, A. Prandini, L. Toti, van den Born G.J., and A. Vespermann, 2009: Climate change and food safety: An emerging issue with special focus on Europe. Food and Chemical Toxicology, 47(5), 1009-1021. Miranda, A.I., E. Marchi, M. Ferretti, and M.M. Millan, 2009: Forest res and air quality issues in Southern Europe. In: Developments in Environmental Science. [Bytnerowicz, A., M. Arbaugh, A. Riebau, and C. Andersen (eds.)]. Elsevier, Amsterdam, pp. 209-231. Mirasgedis, S., Y. Sarafidis, E. Georgopoulou, V. Kotroni, K. Lagouvardos, and D.P. Lalas, 2007: Modelling framework for estimating impacts of climate change on electricity demand at regional level: Case of Greece. Energy Conversion and Management, 48(5), 1737-1750. Mirasgedis, S., E. Georgopoulou, Y. Sarafidis, K. Papagiannaki, and D.P. Lalas, 2013: The impact of climate change on the demand pattern of bottled water and non-alcoholic beverages . Business Strategy and the Environment, (doi: 10.1002/bse.1782). Mitchell, T.D., T.R. Carter, P.D. Jones, M. Hulme, and M. New, 2004: A Comprehensive Set of High-Resolution Grids of Monthly Climate for Europe and the Globe: The Observed Record (1901-2000) and 16 scenarios (2001-2100). Tyndall Centre for Climate Change Research, University of East Anglia, Norwich, UK, Tyndall Centre Working Paper 55 ed., pp. 1-30. Moen, J. and P. Fredman, 2007: Effects of climate change on alpine skiing in Sweden. Journal of Sustainable Tourism, 15(4), 418-437. Mokrech, M., R. Nicholls, J. Richards, C. Henriques, I. Holman, and S. Shackley, 2008: Regional impact assessment of flooding under future climate and socio-economic scenarios for East Anglia and North West England. Climatic Change, 90(1), 31-55. Moller, A.P., D. Rubolini, and E. Lehikoinen, 2008: Populations of migratory bird species that did not show a phenological response to climate change are declining. Proceedings of the National Academy of Sciences of the United States of America, 105(42), 16195-16200. Molnar, J.L., R.L. Gamboa, C. Revenga, and M.D. Spalding, 2008: Assessing the global threat of invasive species to marine biodiversity. Frontiers in Ecology and the Environment, 6(9), 485-492. Montoya, J.M. and D. Raffaelli, 2010: Climate change, biotic interactions and ecosystem services. Philosophical Transactions of the Royal Society of London.Series B, Biological Sciences, 365(1549), 2013-8. Mooij de, R. and P. Tang, 2003: Four Futures of Europe . Centraal Planbureau, Netherlands, pp. 1-220. Mooij, W.M., J.H. Janse, L.N. De Senerpont Domis, S. Hülsmann, and B.W. Ibelings, 2007: Predicting the effect of climate change on temperate shallow lakes with the ecosystem model PCLake . Hydrobiologica, 584(1), 443- 454. Morán-López, R., J.L. Pérez-Bote, E. da Silva, and A.B.P. Casildo, 2012: Hierarchical large-scale to local-scale influence of abiotic factors in summer-fragmented Mediterranean rivers: Structuring effects on fish distributions, assemblage composition and species richness. Hydrobiologia, 696(1), 137-158. Moreira, F., O. Viedma, M. Arianoutsou, T. Curt, N. Koutsias, E. Rigolot, A. Barbati, P. Corona, P. Vaz, G. Xanthopoulos, F. Mouillot, and E. Bilgili, 2011: Landscape - wildfire interactions in southern Europe: Implications for landscape management. Journal of Environmental Management, 92(10), 2389-2402. Moreno, A. and B. Amelung, 2009: Climate change and tourist comfort on Europe s Beaches in summer: a reassessment. Coastal Management, 37(6), 550-568. Moreno, A., 2010: Mediterranean tourism and climate (change): a survey-based study. Tourism and Hospitality Planning & Development, 7(3), 253-265. Moriondo, M., M. Bindi, C. Fagarazzi, R. Ferrise, and G. Trombi, 2011: Framework for high-resolution climate change impact assessment on grapevines at a regional scale. Regional Environmental Change, 11(3), 553-567. Moriondo, M., M. Bindi, Z.W. Kundzewicz, M. Szwed, A. Chorynski, P. Matczak, M. Radziejewski, D. McEvoy, and A. Wreford, 2010: Impact and adaptation opportunities for European agriculture in response to climatic change and variability. Mitigation and Adaptation Strategies for Global Change, 15(7), 657-679. Moriondo, M., C. Pacini, G. Trombi, C. Vazzana, and M. Bindi, 2010b: Sustainability of dairy farming system in Tuscany in a changing climate. European Journal of Agronomy, 32(1), 80-90. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mouillot, D., D.R. Bellwood, C. Baraloto, J. Chave, R. Galzin, M. Harmelin-Vivien, M. Kulbicki, S. Lavergne, S. Lavorel, N. Mouquet, C.E. Paine, J. Renaud, and W. Thuiller, 2013: Rare species support vulnerable functions in high-diversity ecosystems. PLoS Biology, 11(5), e1001569. Mullan, D., D. Favis-mortlock, and R. Fealy, 2012: Agricultural and forest meteorology addressing key limitations associated with modelling soil erosion under the impacts of future climate change. Agricultural and Forest Meteorology, 156, 18-30. Musshoff, O., M. Odening, and W. Xu, 2011: Management of climate risks in agriculture - will weather derivatives permeate? Applied Economics, 43(9), 1067-1077. Mustonen, T. and K. Mustonen, 2011a: Eastern Sámi Atlas. Snowchange Cooperative, Finland, pp. 334. Mustonen, T. and K. Mustonen, 2011b: Drowning Reindeers, Drowning Homes - Indigenous Saami and Hydroelectricity in Sompio, Finland. Snowchange Cooperative, Finland, pp. 116. Nabuurs, G.-., M. Lindner, P.J. Verkerk, K. Gunia, P. Deda, R. Michalak, and G. Grassi, 2013: First signs of carbon sink saturation in European forest biomass. Nature Climate Change, 3(9), 792-796. Najac, J., C. Lac, and L. Terray, 2011: Impact of climate change on surface winds in France using a statistical- dynamical downscaling method with mesoscale modelling. International Journal of Climatology, 31(3), 415- 430. Navarra, A., L.Tubiana (eds.), 2013: Regional Assessment of Climate Change in the Mediterranean. Volume 3: Case Studies., Springer Verlag. ed., pp. 225. Nicholls, R., P. Wong, V. Burkett, C. Woodroffe, and J. Hay, 2008: Climate change and coastal vulnerability assessment: scenarios for integrated assessment. Sustainability Science, 3(1), 89-102. Nicholls, S. and B. Amelung, 2008: Climate change and tourism in Northwestern Europe: impacts and adaptation. Tourism Analysis, 13(1), 21-31. Nokkala, M., P. Leviakangas, and K. Oiva (eds.), 2012: The Costs of Extreme Weather for the European Transport System. EWENT Project D4. Vtt Technology, 36 VTT Technical Research Centre of Finland, Espoo, Finland, pp. 92. Nolan, P., P. Lynch, R. Mcgrath, T. Semmler, and S. Wang, 2012: Simulating climate change and its effect on the the wind energy resource of Ireland. Wind Energy, 15, 593-608. OECD, 2007: Climate Change in the European Alps: Adapting Winter Tourism and Natural Hazards Management. Organisation for Economic Co-operation and Development, Paris, France, pp. 1-136. Okruszko, T., H. Duel, M. Acreman, M. Grygoruk, M. Flörke, and C. Schneider, 2011: Broad-scale ecosystem services of European wetlands overview of the current situation and future perspectives under different climate and water management scenarios. Hydrological Sciences Journal, 56(8), 1501-1517. Olesen, J.E., M. Trnka, K.C. Kersebaum, A.O. Skjelvag, B. Seguin, P. Peltonen-Sainio, F. Rossi, J. Kozyra, and F. Micale, 2011: Impacts and adaptation of European crop production systems to climate change. European Journal of Agronomy, 34(2), 96-112. Oliver, R.J., J.W. Finch, and G. Taylor, 2009: Second generation bioenergy crops and climate change: a review of the effects of elevated atmospheric CO2 and drought on water use and the implications for yield. GCB Bioenergy, 1(2), 97-114. Olonscheck, M., A. Holsten, and J. Kropp, 2011: Heating and cooling demand and related emissions of the German residential building stock under climate change. Energy Policy, 39, 4795-4806. Oort, P.A.J., B.G.H. Timmermans, and A.C.P.M. van Swaaij, 2012: Why farmers' sowing dates hardly change when temperature rises. European Journal of Agronomy, 40, 102-111. OSPAR, 2010: Chapter 12: Regional summaries. In: Quality Status Report 2010. OSPAR Commission, London, pp. 150-161. Pahl-Wostl, C., 2007: Transitions towards adaptive management of water facing climate and global change. Water Resources Management, 21(1), 49-62. Paiva, R.C.D., W. Collischonn, E.B.C. Schettini, J. Vidal, F. Hendrickx, and A. Lopez, 2011: The case studies. In: Modelling The Impact of Climate Change on Water Resources. [Fung, F., A. Lopez, and M.(.). New (eds.)]. John Wiley-Blackwell, Chichester, United Kingdom, pp. 203. Palahi, M., R. Mavsar, C. Gracia, and Y. Birot, 2008: Mediterranean Forests under Focus . International Forestry Review, 10(4), 676-688. Palin, E., H.E. Thornton, C.T. Mathison, R.E. McCarthy, R.T. Clark, and J. Dora, 2013: Future projections of temperature-related climate change impacts on the railway network of Great Britain. Climatic Change, 120(1- 2), 71-93. Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pall, P., T. Aina, D.A. Stone, P.A. Stott, T. Nozawa, A.G.J. Hilberts, D. Lohmann, and M.R. Allen, 2011: Anthropogenic greenhouse gas contribution to flood risk in England and Wales in autumn 2000. Nature, 470 (7334), 382-385. Paranjothy, S., J. Gallacher, R. Amlôt, G.J. Rubin, L. Page, T. Baxter, J. Wight, D. Kirrage, R. McNaught, and S.R. Palmer, 2011: Psychosocial impact of the summer 2007 flood in England. BMC Public Health, 11(145), 1-8. Parent, B. and F. Tardieu, 2012: Temperature responses of developmental processes have not been affected by breeding in different ecological areas for 17 crop species. New Phytologist, 194(3), 760-774. Pa¹ièko, R., È. Brankoviæ, and Z. Simic, 2012: Assessment of climate change impacts in energy generation from renewable sources in Croatia. Renewable Energy, 46, 224-231. Paterson, J.S., M.B. Araújo, P.M. Berry, J.M. Piper, and M.D.A. Rounsevell, 2008: Mitigation, adaptation, and the threat to biodiversity. Conservation Biology, 22(5), 1352-1355. Paterson, R.R.M. and N. Lima, 2010: How will climate change affect mycotoxins in food? Food Research International, 43(7), 1902-1914. Pauli, H., M. Gottfried, S. Dullinger, O. Abdaladze, M. Akhalkatsi, J.L. Benito Alonso, G. Coldea, J. Dick, B. Erschbamer, R. Fernández Calzado, D. Ghosn, J.I. Holten, R. Kanka, G. Kazakis, J. Kollár, P. Larsson, P. Moiseev, D. Moiseev, U. Molau, J. Molero Mesa, L. Nagy, G. Pelino, M. Puºcaº, G. Rossi, A. Stanisci, A.O. Syverhuset, J.-. Theurillat, M. Tomaselli, P. Unterluggauer, L. Villar, P. Vittoz, and G. Grabherr, 2012: Recent Plant Diversity Changes on Europe s Mountain Summits . Science, 336, 353-355. Pausas, J.G., J. Llovet, A. Rodrigo, and R. Vallejo, 2008: Are wildfires a disaster in the Mediterranean basin? A review. International Journal of Wildland Fire, 17(6), 713-723. Pausas, J.G. and S. Fernández-Munoz, 2012: Fire regime changes in the Western Mediterranean Basin: From fuel- limited to drought-driven fire regime. Climatic Change, 110(1-2), 215-226. Pejovic, T., V.A. Williams, R.B. Noland, and R. Toumi, 2009: Factors affecting the frequency and severity of airport weather delays and the implications of Climate Change for future delays. Transportation Research Record, 2139, 97-106. Pellizzaro, G., A. Ventura, B. Arca, A. Arca, P. Duce, V. Bacciu, and D. Spano, 2010: Estimating effects of future climate on duration of fire danger season in Sardinia[Viegas, D.X. (ed.)]. Proceedings of VI International Forest Fire Research Conference, 15-18, November 2010, Coimbra, Portugal, pp. 123. Peltonen-Sainio, P., L. Jauhiainen, and I.P. Laurila, 2009: Cereal yield trends in northern European conditions: Changes in yield potential and its realisation. Field Crops Research, 110(1), 85-90. Peltonen-sainio, P., L. Jauhiainen, and K. Hakala, 2010: Crop responses to temperature and precipitation according to long-term multi-location trials at high-latitude conditions. The Journal of Agricultural Science, 149(01), 49- 62. Perch-Nielsen, S.L., B. Amelung, and R. Knutti, 2010: Future climate resources for tourism in Europe based on the daily Tourism Climatic Index. Climatic Change, 103(3-4), 363-381. Pereira, M., R. Trigo, C. da Camara, J. Pereira, and S. Leite, 2005: Synoptic patterns associated with large summer forest fires in Portugal. Agricultural and Forest Meteorology, 129(1-2), 11-25. Perez, F.F., X.A. Padin, Y. Pazos, M. Gilcoto, M. Cabanas, P.C. Pardo, M.D. Doval, and L. Farina-Busto, 2010: Plankton response to weakening of the Iberian coastal upwelling. Global Change Biology, 16(4), 1258-1267. Perry, I.A., R. Ommer, K. Cochrane, and P. Cury, 2011: World Fisheries. Wiley-Blackwell, Oxford, pp. 148. Perry, R.I., R.E. Ommer, M. Barange, and F. Werner, 2010: The challenge of adapting marine social-ecological systems to the additional stress of climate change. Current Opinion in Environmental Sustainability, 2(5-6), 356-363. Peterson, T., P. Scott, and S. Herring, 2012: Explaining extreme events of 2011 from a climate perspective. Bulletin of the American Meteorological Society, 93, 1041-1067. Petitpierre, B., C. Kueffer, O. Broennimann, C. Randin, C. Daehler, and A. Guisan, 2012: Climatic niche shifts are rare among terrestrial plant invaders. Science, 335(6074), 1344-1348. Petney, T.N., J. Skuballa, S. Muders, M. Pfäffle, C. Zetlmeisl, and R. Oehme, 2012: The Changing Distribution Patterns of Ticks (Ixodida) in Europe in Relation to Emerging Tick-Borne Diseases. Parasitology Research Monographs, 3, 151-166. Petrow, T., B. Merz, K.E. Lindenschmidt, and A.H. Thieken, 2007: Aspects of seasonality and flood generating circulation patterns in a mountainous catchment in south-eastern Germany. Hydrology and Earth System Sciences, 11, 1455-1468. Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Petrow, T., J. Zimmer, and B. Merz, 2009: Changes in the flood hazard in Germany through changing frequency and persistence of circulation patterns. Natural Hazards and Earth System Sciences, 9(4), 1409-1423. Pfenniger, S., S. Hanger, and M.a.a. Dreyfus, 2010: Report on perceived policy needs and decision contexts, Mediation Deliverable 1.1 (Final Draft). subject to approval by the European Commission, pp. 108. Philippart, C.J.M., R. Anadon, R. Danovaro, J.W. Dipper, K.F. Drinkwater, S.J. Hawkins, T. Oguz, G. O Sullivan, and P.C. Reid, 2011: Impacts of climate change on European marine ecosystems: observations, expectations and indicators. Journal of Experimental Marine Biology and Ecology, 400, 52-69. Pilli-Sihlova, K., P. Aatola, M. Ollikainen, and H. Tuomenvirta, 2010: Climate Change and electricity consumption - witnessing increasing or decreasing use and costs? Energy Policy, 38(5), 2409-2419. Pinto, J.G., U. Ulbrich, G.C. Leckebusch, T. Spangehl, M. Reyers, and S. Zacharias, 2007a: Changes in storm track and cyclone activity in three SRES ensemble experiments with the ECHAM5/MPI-OM1 GCM. Climate Dynamics, 29(2-3), 195-210. Pinto, J.G., C.P. Neuhaus, G.C. Leckebusch, M. Reyers, and M. Kerschgens, 2010: Estimation of wind storm impacts over Western Germany under future climate conditions using a statistical-dynamical downscaling approach. Tellus, 62(2), 188-201. Pinto, J.G., E.L. Fröhlich, G.C. Leckebusch, and U. Ulbrich, 2007b: Changing European storm loss potentials under modified climate conditions according to ensemble simulations of the ECHAM5/MPI-OM1 GCM. Natural Hazards and Earth System Sciences, 7(1), 165-175. Pitois, S.G. and C.J. Fox, 2006: Long-term changes in zooplankton biomass concentration and mean size over the Northwest European shelf inferred from Continuous Plankton Recorder data. ICES Journal of Marine Science, 63(5), 785-798. Pitt, M., 2008: The Pitt Review: Lessons Learned from the 2007 floods. Final Report. Cabinet Office, London, UK, pp. 1-505. Planque, B., J. Fromentin, P. Cury, K.F. Drinkwater, S. Jennings, R.I. Perry, and S. Kifani, 2010: How does fishing alter marine populations and ecosystems sensitivity to climate? Journal of Marine Systems, 79(3-4), 403-417. Planton, S., P. Lionello, V. Artale, R. Aznar, A. Carillo, J. Colin, L. Congedi, C. Dubois, A. Elizalde Arellano, S. Gualdi, E. Hertig, G. Jorda Sanchez, L. Li, J. Jucundus, C. Piani, P. Ruti, E. Sanchez-Gomez, G. Sannino, F. Sevault, and S. Somot, 2011: The climate of the Mediterranean region under global warming. In: Mediterranean Climate Variability. [Lionello, P., P. Malanotte-Rizzoli, and R.(.). Boscolo (eds.)]. Elsevier, Netherlands, pp. 399-416. Poirier, M., J.L. Durand, and F. Volaire, 2012: Persistence and production of perennial grasses under water deficits and extreme temperatures: importance of intraspecific vs. interspecific variability. Global Change Biology, 18(12), 3632-3646. Polemio, M. and O. Petrucci, 2010: Occurrence of landslide events and the role of climate in the twentieth century in Calabria, southern Italy. Quarterly Journal of Engineering Geology and Hydrogeology, 43(4), 403-415. Popov Janevska, D., R. Gospavic, E. Pacholewicz, and V. Popov, 2010: Application of HACCP-QMRA approach for managing the impact of climate change on food quality and safety. Food Research International, 43(7), 1915-1924. Post, J., T. Conradt, F. Suckow, V. Krysanova, F. Wechsung, and F.F. Hattermann, 2008: Integrated assessment of cropland soil carbon sensitivity to recent and future climate in the Elbe River basin. Hydrological Sciences Journal, 53(5), 1043-1058. Powlson, D.S., A.P. Whitmore, and K.W.T. Goulding, 2011: Soil carbon sequestration to mitigate climate change: A critical re-examination to identify the true and the false. European Journal of Soil Science, 62(1), 42-55. Pruszak, Z. and E. Zawadzka, 2008: Potential implications of sea-level rise for Poland. Journal of Coastal Research, 24(2), 410-422. Pryor, S.C. and R.J. Barthelmie, 2010: Climate change impacts on wind energy: a review. Renewable and Sustainable Energy Reviews, 14(1), 430-437. Pryor, S.C. and J.T. Schoof, 2010: Importance of the SRES in projections of climate change impacts on near-surface wind regimes. Meteorologische Zeitschrift, 19(3), 267-274. Purvis, M.J., P.D. Bates, and C.M. Hayes, 2008: A probabilistic methodology to estimate future coastal flood risk due to sea level rise. Coastal Engineering, 55(12), 1062-1073. Quevauviller, P., 2011: Adapting to climate change: reducing water-related risks in Europe EU policy and research considerations. Environmental Science & Policy, 14(7), 722-729. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Queyriaux, B., A. Armengaud, C. Jeannin, C. Coutourier, and F. Peloux-Petiot, 2008: Chikungunya in Europe. The Lancet, 371(9614), 723-724. Quintana-Segui, P., F. Habets, and E. Martin, 2011: Comparison of past and future Mediterranean high and low extremes of precipitation and river flow projected using different statistical downscaling methods. Natural Hazards and Earth System Sciences, 11(5), 1411-1432. Radovic, V., K. Vitale, and P.B. Tchounwou, 2012: Health facilities safety in natural disasters: experiences and challenges from South East Europe. International Journal of Environmental Research and Public Health, 9(5), 1677-1686. Raftoyannis, Y., I. Spanos, and K. Radoglou, 2008: The decline of Greek fir (Abies cephalonica Loudon): relationships with root condition. Plant Biosystems, 142(2), 386-390. Rahel, F.J. and J.D. Olden, 2008: Assessing the effects of climate change on aquatic invasive species. Conservation Biology, 22(13), 521-533. Räisänen, J. and J. Eklund, 2012: 21st Century changes in snow climate in Northern Europe: a high-resolution view from ENSEMBLES regional climate models. Climate Dynamics, 38(11-12), 2575-2591. Randolph, S.E. and D.J. Rogers, 2010: The arrival, establishment and spread of exotic diseases: patterns and predictions. Nature Reviews, 8(5), 361-371. Rauthe, M., M. Kunz, and C. Kottmeier, 2010: Changes in wind gust extremes over Central Europe derived from a small ensemble of high resolution regional climate models. Meteorologische Zeitschrift, 19(3), 299-312. Ready, P.D., 2010: Leishmaniasis emergence in Europe. Eurosurveillance, 15(10), 29-39. Rees, P., N. van der Gaag, J. de Beer, and F. Heins, 2012: European regional populations: current trends, future pathways, and policy options. European Journal of Population, 28(4), 385-416. Rees, W.G., F.M. Stammler, F.S. Danks, and P. Vitebsky, 2008: Vulnerability of European reindeer husbandry to global change. Climatic Change, 87(1-2), 199-217. Refsgaard, J.C., K. Arnbjerg-Nielsen, M. Drews, K. Halsnaes, E. Jeppesen, H. Madsen, A. Markandya, J.E. Olesen, J.R. Porter, and J.H. Christensen, 2013: The role of uncertainty in climate change adaptation strategies. A Danish water management example. Mitigation and Adaptation Strategies for Global Change, 18(3), 337-359. Reginster, I. and M. Rounsevell, 2006: Scenarios of future urban land use in Europe. Environment and Planning B: Planning & Design, 33(4), 619-636. Renard, B., M. Lang, P. Bois, A. Dupeyrat, O. Mestre, H. Niel, E. Sauquet, C. Prudhomme, S. Parey, E. Paquet, L. Neppel, and J. Gailhard, 2008: Regional methods for trend detection: Assessing field significance and regional consistency. Water Resources Research, 44(8), W08419. Renaudeau, D., J.L. Gourdine, and N.R. St-Pierre, 2011: Meta-analysis of the effect of high ambient temperature on growing finishing pigs. Journal of Animal Science, 89(7), 2220-2230. Renaudeau, D., A. Collin, S. Yahav, V. De Basilio, J.L. Gourdine, and R.J. Collier, 2012: Adaptation to hot climate and strategies to alleviate heat stress in livestock production. Animal, 6(5), 707-728. Resco, d.D., C. Fischer, and C. Colinas, 2007: Climate Change Effects on Mediterranean Forests and Preventive Measures. New Forests, 33(1), 29-40. Revich, B., 2011: Heat-wave, air quality and mortality in the Russian Federation's Europe, 2010: Preliminary assessment. Human Ecology, 7, 3-9. Revich, B. and D.A. Shaposhnikov, 2012: Climate Change, Heat Waves, and Cold Spells as Risk Factors for Increased Mortality in Some Regions of Russia. Studies on Russian Economic Development, 23(2), 195-207. Rickebusch, S., W. Thuiller, T. Hickler, M.B. Araújo, M.T. Sykes, O. Schweiger, and B. Lafourcade, 2008: Incorporating the effects of changes in vegetation functioning and CO2 on water availability in plant habitat models. Biology Letters, 4(5), 556-559. Rico-Amoros, A.M., J. Olcina-Cantosa, and D. Sauri, 2009: Tourist land use patterns and water demand: evidence from the Western Mediterranean. Land use Policy, 26(2), 493-501. Rigling, A., C. Bigler, B. Eilmann, E. Feldmeyer-Christe, U. Gimmi, C. Ginzler, U. Graf, P. Mayer, G. Vacchiano, P. Weber, T. Wohlgemuth, R. Zweifel, and M. Dobbertin, 2013: Driving factors of a vegetation shift from Scots pine to pubescent oak in dry Alpine forests. Global Change Biology, 19, 229-240. Rixen, C., M. Teich, C. Lardelli, D. Gallati, M. Pohl, M. Pütz, and P. Bebi, 2011: Winter tourism and climate change in the Alps: an assessment of resource consumption, snow reliability, and future snowmaking potential. Mountain Research and Development, 31(3), 229-236. Robine, J.M., S.L.K. Cheung, S. Le Roy, H. Van Oyen, C. Griffiths, J.P. Michel, and F.R. Herrmann, 2008: Death toll exceeded 70,000 in Europe during the summer of 2003. Comptes Rendus - Biologies, 331(2), 171-178. Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rockel, B. and K. Woth, 2007: Extremes of near-surface wind speed over Europe and their future changes as estimated from an ensemble of RCM simulations. Climatic Change, 81(Suppl 1), 267-280. Rocklöv, J. and B. Forsberg, 2010: The effect of high ambient temperature on the elderly population in three regions of Sweden. International Journal of Environmental Research and Public Health, 7(6), 2607-2619. Roiz, D., M. Neteler, C. Castellani, D. Arnoldi, and A. Rizzoli, 2011: Climatic factors driving invasion of the tiger mosquito (Aedes albopictus) into new areas of Trentino, northern Italy. PLoS One, 6(4), art. no. e14800. Rojas, R., L. Feyen, and P. Watkiss, 2013: Climate change and river floods in the European Union: socio-economic consequences and the costs and benefits of adaptation. Global Environmental Change, 10.1016/j.gloenvcha.2013.08.006. Rojas, R., L. Feyen, A. Bianchi, and A. Dosio, 2012: Assessment of future flood hazard in Europe using a large ensemble of bias-corrected regional climate simulations. Journal of Geophysical Research D: Atmospheres, 117(17), art. no. D17109. Roos, J., R. Hopkins, A. Kvarnheden, and C. Dixelius, 2011: The impact of global warming on plant diseases and insect vectors in Sweden. European Journal of Plant Pathology, 129(1), 9-19. Rosan, P. and D. Hammarlund, 2007: Effects of climate, fire and vegetation development on Holocene changes in total organic carbon concentration in three boreal forest lakes in northern Sweden. Biogeosciences, 4(6), 975- 984. Rosenzweig, C., D. Karoly, M. Vicarelli, P. Neofotis, Q. Wu, G. Casassa, A. Menzel, T.L. Root, N. Estrella, B. Seguin, P. Tryjanowski, C. Liu, S. Rawlins, and A. Imeson, 2008: Attributing physical and biological impacts to anthropogenic climate change. Nature, 453, 353-357. Rosenzweig, C. and F.N. Tubiello, 2007: Adaptation and mitigation strategies in agriculture: an analysis of potential synergies. Mitigation and Adaptation Strategies for Global Change, 12(5), 855-873. Rötter, R.P., T. Palosuo, N.K. Pirttioja, M. Dubrovski, T. Salo, S. Fronzek, R. Aikasalo, M. Trnka, A. Ristolainen, and T. Carter, 2011: What would happen to barley production in Finland if global warming exceeded 4 °C? A model-based assessment. European Journal of Agronomy, 35(4), 205-214. Rouault, G., J.N. Candau, F. Lieutier, L.M. Nageleisen, J.C. Martin, and N. Warzée, 2006: Effects of drought and heat on forest insect populations in relation to the 2003 drought in Western Europe. Annals of Forest Science, 63(6), 613-624. Rounsevell, M.D.A., I. Reginster, M.B. Araújo, T.R. Carter, N. Dendoncker, F. Ewert, J.I. House, S. Kankaanpää, R. Leemans, M.J. Metzger, C. Schmit, P. Smith, and G. Tuck, 2006: A coherent set of future land use change scenarios for Europe. Agriculture, Ecosystems and Environment, 114(1), 57-68. Rounsevell, M.D.A. and D.S. Reay, 2009: Land use and climate change in the UK. Land use Policy, 26(Suppl 1), 160-169. Rounsevell, M.D.A. and M.J. Metzger, 2010: Developing qualitative scenario storylines for environmental change assessment. Wiley Interdisciplinary Reviews: Climate Change, 1(4), 606-619. Rubolini, D., R. Ambrosini, M. Caffi, P. Brichetti, S. Armiraglio, and N. Saino, 2007a: Long-term trends in first arrival and first egg laying dates of some migrant and resident bird species in northern Italy. International Journal of Biometeorology, 51(6), 553-563. Rubolini, D., A.P. Mller, K. Rainio, and E. Lehikoinen, 2007b: Intraspecific consistency and geographic variability in temporal trends of spring migration phenology among European bird species. Climate Research, 35(1-2), 135-146. Ruiz-Ramos, D.V., E.A. Hernandez-Delgado, and N.V. Schizas, 2011: Population status of the long-spined urchin Diadema antillarum in Puerto Rico 20 years after a mass mortality event. Bulletin of Marine Science, 87(1), 113-127. Rusch, G.M., 2012: Climate and ecosystem services. The potential of Norwegian ecosystems for climate mitigation and adaptation. In: NINA Report 791. Norwegian Institute for Nature Research, Trondheim, pp. 1-43. Rutty, M. and D. Scott, 2010: Will the Mediterranean become too hot for tourism? A reassessment. Tourism Planning & Development, 7(3), 267-281. Sabbioni, C., A. Bonazza, and P. Messina, 2008: Global climate change and archaeological heritage: prevision, impact and mapping. In: ARCHAIA. Case Studies on Research Planning, Characterisation, Conservation and Management of Archaeological Sites . [Marchetti, N. and I. Thuesen (eds.)]. Archaeopress, Oxford, pp. 295- 300. Sabbioni, C., P. Brimblecombe, and M. Cassar, 2012: Atlas of Climate Change Impact on European Cultural Heritage. Anthem Press, London, pp. 160. Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sabir, M., J. Ommeren, M. Koetse, and P. Rietveld, 2010: Adverse weather and commuting speed. Networks and Spatial Economics, 11(4), 701-712. Saino, N., R. Ambrosini, D. Rubolini, J. Von Hardenberg, A. Provenzale, K. Hüppop, O. Hüppop, A. Lehikoinen, E. Lehikoinen, K. Rainio, M. Romano, and L. Sokolov, 2011: Climate warming, ecological mismatch at arrival and population decline in migratory birds. Proceedings of the Royal Society B: Biological Sciences, 278(1707), 835-842. Sainz-Elipe, S., J.M. Latorre, R. Escosa, M. Masia, M.V. Fuentes, S. Mas-Coma, and M.D. Bargues, 2010: Malaria resurgence risk in Southern Europe: climate assessment in an historically endemic area of rice fields at the Mediterranean shore of Spain. Malaria Journal, 9(1), 221-237. Salis, M., A. Ager, B. Arca, M. Finney, V. Bacciu, P. Duce, and D. Spano, 2013: Assessing exposure of human and ecological values to wildfire in Sardinia, Italy. International Journal of Wildland Fire, 22(4), 549-565. Sanchez-Rodriguez, R., 2009: Learning to adapt to climate change in urban areas. A review of recent contributions. Current Opinion in Environmental Sustainability, 1(2), 201-206. San-Miguel-Ayanz, J., M. Rodrigues, S. Santos de Oliveira, C.K. Pacheco, F. Moreira, B. Duguy, and A. Camia, 2012: Land Cover Change and Fire Regime in the European Mediterranean Region. In: Post-Fire Management and Restoration of Southern European Forests. Managing Forest Ecosystems. [Moreira, F., M. Arianoutsou, P. Corona, and J. De las Heras (eds.)]. Springer, Netherlands, pp. 21-43. San-Miguel-Ayanz, J., J.M. Moreno, and A. Camia, 2013: Analysis of large fires in European Mediterranean landscapes: Lessons learned and perspectives. Forest Ecology and Management, 294, 11-22. Santos, J.A., A.C. Malheiro, M.K. Karremann, and J.G. Pinto, 2011: Statistical modelling of grapevine yield in the Port Wine region under present and future climate conditions. International Journal of Biometeorology, 55(2), 119-131. Sauter, T., C. Weitzenkamp, and C. Schneider, 2010: Spatio-temporal prediction of snow cover in the Black Forest mountain range using remote sensing and a recurrent neural network. International Journal of Climatology, 30(15), 2330-2341. Savé, R., F. de Herralde, X. Aranda, E. Pla, D. Pascual, I. Funes, and C. Biel, 2012: Potential changes in irrigation requirements and phenology of maize, apple trees and alfalfa under global change conditions in Fluvia watershed during XXIst century: Results from a modeling approximation to watershed-level water balance. Agricultural Water Management, 114, 78-87. Schaefli, B., B. Hingray, and A. Musy, 2007: Climate change and hydropower production in the Swiss Alps: quantification of potential impacts and related modeling uncertainties. Hydrology & Earth System Sciences, 11(3), 1191-1205. Schär, C. and G. Jendritzky, 2004: Climate change: hot news from summer 2003. Nature, 432(7017), 559-560. Schifano, P., M. Leone, M. De Sario, F. de'Donato, A.M. Bargagli, D. D'Ippoliti, C. Marino, and P. Michelozzi, 2012: Changes in the effects of heat on mortality among the elderly from 1998-2010: results from a multicentre time series study in Italy. Environmental Health: A Global Access Science Source, 11(1), Article number58. Schmidli, J., C.M. Goodess, C. Frei, M.R. Haylock, Y. Hundecha, J. Ribalaygua, and T. Schmith, 2007: Statistical and dynamical downscaling of precipitation: an evaluation and comparison of scenarios for the European Alps. Journal of Geophysical Research, 112(4), Article numberD04105. Schmocker-Fackel, P. and F. Naef, 2010: Changes in flood frequencies in Switzerland since 1500. Hydrology and Earth System Sciences, 14(8), 1581-1594. Schnitzler, J., J. Benzler, D. Altmann, I. Mucke, and G. Krause, 2007: Survey on the population's needs and the public health response during floods in Germany 2002. Journal of Public Health Management and Practice, 13(5), 461-464. Scholz, G., J.N. Quinton, and P. Strauss, 2008: Soil erosion from sugar beet in Central Europe in response to climate change induced seasonal precipitation variations. Catena, 72(1), 91-105. Schroter, D., W. Cramer, R. Leemans, I. Prentice, M. Araujo, N. Arnell, A. Bondeau, H. Bugmann, T. Carter, A. Vega-Leinert, M. Erhard, F. Ewert, M. Glendining, J. House, S. Kankaanpaa, R. Klein, S. Lavorel, M. Lindner, M. Metzger, J. Meyer, T. Mitchell, I. Reginster, M. Rounsevell, S. Sabate, S. Sitch, B. Smith, J. Smith, P. Smith, M. Sykes, K. Thonicke, W. Thuiller, G. Tuck, S. Saehle, and B. Zierl, 2005: Ecosystem Service Supply and Vulnerability to Global Change in Europe. Science, 310, 1333-1337. Schulze, E.D., S. Luyssaert, P. Ciais, A. Freibauer, I.A. Janssens, J.F. Soussana, P. Smith, J. Grace, I. Levin, B. Thiruchittampalam, M. Heimann, A.J. Dolman, R. Valentini, P. Bousquet, P. Peylin, W. Peters, C. Rödenbeck, G. Etiope, N. Vuichard, M. Wattenbach, G.J. Nabuurs, Z. Poussi, J. Nieschulze, and J.H. Gash, 2010: Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Importance of methane and nitrous oxide for Europe's terrestrial greenhouse-gas balance. Nature Geoscience, 2(12), 842-850. Schutze, N. and G.H. Schmitz, 2010: OCCASION: new planning tool for optimal climate change adaption strategies in irrigation. Journal of Irrigation and Drainage Engineering-Asce, 136(12), 836-846. Schwarze, R., M. Schwindt, H. Weck-Hannemann, P. Raschky, F. Zahn, and G. Wagner, 2011: Natural Hazard Insurance in Europe: Tailored Responses to Climate Change are Needed. Environmental Policy and Governance, 21, 14-30. Schweiger, O., R.K. Heikkinen, A. Harpke, T. Hickler, S. Klotz, O. Kudrna, I. Kühn, J. Pöyry, and J. Settele, 2012: Increasing range mismatching of interacting species under global change is related to their ecological characteristics. Global Ecology and Biogeography, 21(1), 88-99. Schwierz, C., P. Köllner-Heck, E.Z. Mutter, D.N. Bresch, P.L. Vidale, M. Wild, and C. Schär, 2010: Modelling European winter wind storm losses in current and future climate. Climatic Change, 101(3), 485-514. Seidl, R., M.-. Schelhaas, M. Lindner, and M.J. Lexer, 2009: Modelling bark beetle disturbances in a large scale forest scenario model to assess climate change impacts and evaluate adaptive management strategies. Regional Environmental Change, 9(2), 101-119. Seidl, R., W. Rammer, and M. Lexer, 2011: Climate change vulnerability of sustainable forest management in the Eastern Alps. Climatic Change, 106, 225-254. Seidl, R. and M. Lexer, 2013: Forest management under climatic and social uncertainty: Trade-offs between reducing climate change impacts and fostering adaptive capacity. Journal of Environmental Management, 114, 461-469. Seljom, P., E. Rosenberg, A. Fidge, J. Haugen, M. Meir, J. Rekstad, and T. Jarlset, 2011: Modelling the effects of climate change on the energy system-A case study of Norway. Energy Policy, 39(11), 7310-7321. Semenov, M.A., 2009: Impacts of climate change on wheat in England and Wales. Journal of the Royal Society Interface, 6(33), 343-350. Semenov, V.A., 2011: Climate-related changes in hazardous and adverse hydrological events in the Russian rivers. Russian Meteorology and Hydrology, 36(2), 124-129. Semenza, J.C. and B. Menne, 2009: Climate change and infectious diseases in Europe. Lancet Infectious Diseases, 9(6), 365-375. Semenza, J., J. Suk, V. Estevez, K.L. Ebi, and E. Lindgren, 2012: Mapping climate change vulnerabilities to infectious diseases in Europe. Environmental Health Perspectives, 120(3), 385-392. Senatore, A., G. Mendicino, G. Smiatek, and H. Kunstmann, 2011: Regional climate change projections and hydrological impact analysis for a Mediterranean basin in Southern Italy. Journal of Hydrology, 399(1-2), 70- 92. Serquet, G. and M. Rebetez, 2011: Relationship between tourism demand in the Swiss Alps and hot summer air temperatures associated with climate change. Climatic Change, 108(1), 291-300. Shvidenko, A.Z., D.G. Shchepashchenko, E.A. Vaganov, A.I. Sukhinin, S.S. Maksyutov, I. McCallum, and I.P. Lakyda, 2011: Impact of wildfire in Russia between 1998-2010 on ecosystems and the global carbon budget. Doklady Earth Sciences, 441(2), 1678-1682. Siebert, S. and F. Ewert, 2012: Spatio-temporal patterns of phenological development in Germany in relation to temperature and day length. Agricultural and Forest Meteorology, 152, 44-57. Silva, D.E., P. Rezende Mazzella, M. Legay, E. Corcket, and J.L. Dupouey, 2012: Does natural regeneration determine the limit of European beech distribution under climatic stress? Forest Ecology and Management, 266, 263-272. Sirotenko, O.D. and E.V. Abashina, 2008: Modern Climate Changes of Biosphere Productivity in Russia and Adjacent Countries. Russian Meteorology and Hydrology, 33(4), 267-271. Skeffington, M.S. and K. Hall, 2011: The ecology, distribution and invasiveness of Gunnera L. species in Connemara, Western Ireland. Biology and Environment, 111(3), 157-175. Slangen, A., C. Katsman, R. van de Wal, L. Vermeersen, and R. Riva, 2012: Towards regional projections of twenty-first century sea-level change based on IPCC SRES scenarios. Climate Dynamics, 38, 1191-1209. Smith, P. and J.E. Olesen, 2010: Synergies between mitigation of, and adaptation to, climate change in agriculture. Journal of Agricultural Science, 148(5), 543-552. Smith, T.M., R.W. Reynolds, T.C. Peterson, and J. Lawrimore, 2008: Improvements to NOAA's Historical Merged Land-Ocean Surface Temperature analysis (1880-2006). Journal of Climate, 21(10), 2283-2293. Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Smith, P., P.J. Gregory, D. van Vuuren, M. Obersteiner, P. Havlik, M. Rounsevell, J. Woods, E. Stehfest, and J. Bellarby, 2010: Competition for land. Philosophical Transactions of the Royal Society of London.Series B, Biological Sciences, 365(1554), 2941-2957. Soane, B.D., B.C. Ball, J. Arvidsson, G. Basch, F. Moreno, and J. Roger-Estrade, 2012: No-till in northern, western and south-western Europe: A review of problems and opportunities for crop production and the environment. Soil and Tillage Research, 118, 66-87. Solberg, S., O. Hov, A. Svde, I.S.A. Isaksen, P. Coddeville, H. De Backer, C. Forster, Y. Orsolini, and K. Uhse, 2008: European surface ozone in the extreme summer 2003. Journal of Geophysical Research, 113(7), Article numberD07307 (doi:10.1029/2007JD009098). Sorte, C.J.B., S.L. Williams, and R.A. Zerebecki, 2010: Ocean warming increases threat of invasive species in a marine fouling community. Ecology, 91, 2198-2204. Sousa, P.M., R.M. Trigo, P. Aizpurua, R. Nieto, L. Gimeno, and R. Garcia-Herrera, 2011: Trends and extremes of drought indices throughout the 20th century in the Mediterranean. Natural Hazards and Earth System Science, 11(1), 33-51. Soussana, J.F. and A. Luscher, 2007: Temperate grasslands and global atmospheric change: a review. Grass and Forage Science, 62(2), 127-134. Soussana, J.F., A.I. Graux, and F.N. Tubiello, 2010: Improving the use of modelling for projections of climate change impacts on crops and pastures. Journal of Experimental Botany, 61(8), 2217-2228. Spangenberg, L., F. Battke, M. Grana, K. Nieselt, and H. Naya, 2011: Identifying associations between amino acid changes and meta information in alignments. Bioinformatics, 27(20), 2782-2789. Stafoggia, M., F. Forastiere, D. Agostini, N. Caranci, F. de'Donato, M. Demaria, P. Michelozzi, R. Miglio, M. Rognoni, A. Russo, and C.A. Perucci, 2008: Factors affecting in-hospital heat-related mortality: a multi-city case-crossover analysis. Journal of Epidemiology and Community Health, 62(3), 209-215. Stahl, K., H. Hisdal, J. Hannaford, L.M. Tallaksen, H.A.J. van Lanen, E. Sauquet, S. Demuth, M. Fendekova, and J. Jódar, 2010: Streamflow trends in Europe: evidence from a dataset of near-natural catchments. Hydrology & Earth System Sciences, 14, 2367-2382. Stanzel, P. and H.P. Nachtnebel, 2010: Mögliche Auswirkungen des Klimawandels auf den Wasserhaushalt und die Wasserkraftnutzung in Österreich. Österreichische Wasser-Und Abfallwirtschaft, 62(9-10), 180-187. Steele-Dunne, S., P. Lynch, R. McGrath, T. Semmler, S. Wang, J. Hanafin, and P. Nolan, 2008: The impacts of climate change on hydrology in Ireland. Journal of Hydrology, 356(1-2), 28-45. Steger, C., S. Kotlarski, T. Jonas, and C. Schär, 2013: Alpine snow cover in a changing climate: a regional climate model perspective. Climate Dynamics, 41(3-4), 735-754. Steiger, R. and M. Mayer, 2008: Snowmaking and climate change: future options for snow production in Tyrolean ski resorts. Mountain Research and Development, 28(3-4), 292-298. Steiger, R., 2010: The impact of climate change on ski season length and snowmaking requirements in Tyrol, Austria. Climate Research, 43, 251-262. Steiger, R., 2011: The impact of snow scarcity on ski tourism. An analysis of the record warm season 2006/07 in Tyrol (Austria). Tourism Review, 66(3), 4-13. Steiger, R., 2012: Scenarios for skiing tourism in Austria: integrating demographics with an analysis of climate change. Journal of Sustainable Tourism, 20(6), 867-882. Sterl, A., H. van den Brink, H. de Vries, R. Haarsma, and E. van Meijgaard, 2009: An ensemble study of extreme storm surge related water levels in the North Sea in a changing climate. Ocean Science, 5, 369-378. Stoate, C., A. Báldi, P. Beja, N.D. Boatman, I. Herzon, A. van Doorn, G.R. de Snoo, L. Rakosy, and C. Ramwell, 2009: Ecological impacts of early 21st century agricultural change in Europe--a review. Journal of Environmental Management, 91(1), 22-46. Stoffel, M. and C. Huggel, 2012: Effects of climate change on mass movements in mountain environments. Progress in Physical Geography, 36(3), 421-439. Stoll, S., H.J. Hendricks Franssen, M. Butts, and W. Kinzelbach, 2011: Analysis of the impact of climate change on groundwater related hydrological fluxes: a multi-model approach including different downscaling methods. Hydrology & Earth System Sciences, 15, 21-38. Storm, J., A.W. Cattaneo, and F. Trincardi, 2008: Coastal dynamics under conditions of rapid sea-level rise: Late Pleistocene to Early Holocene evolution of barrier-lagoon systems on the Northern Adriatic shelf (Italy). Quaternary Science Reviews, 27(11-12), 1107-1123. Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Stratonovitch, P., 2012: A process-based approach to modelling impacts of climate change on the damage niche of an agricultural weed. Global Change Biology, 18(6), 2071-2080. Streftaris, N., A. Zenetos, and E. Papathanassiou, 2005: Globalisation in marine ecosystems: the story of non- indigenous marine species across European seas. Oceanogr Mar Biol-an Annual Review, 43, 419-453. Supit, I., C.A. van Diepen, A.J.W. de Wit, P. Kabat, B. Baruth, and F. Ludwig, 2010: Recent changes in the climatic yield potential of various crops in Europe. Agricultural Systems, 103, 683-694. Surminski, S. and A. Philp, 2010: Briefing: guidance on insurance issues for new developments. Proceedings of the Institution of Civil Engineers: Engineering Sustainability, 163(1), 3-6. Swart, R., L. Bernstein, M. Ha-Duong, and A. Petersen, 2009a: Agreeing to disagree: uncertainty management in assessing climate change, impacts and responses by the IPCC. Climatic Change, 92(1), 1-29. Swart, R., R. Biesbroek, S. Binnerup, T.R. Carter, C. Cowan, T. Henrichs, S. Loquen, H. Mela, M. Morecroft, M. Reese, and D. Rey, 2009b: Europe Adapts to Climate Change: Comparing National Adaptation Strategies. PEER Report No. 1. Partnership for European Environmental Research, Helsinki, pp. 283. Swedish Commission on Climate and Vulnerability, 2007: Sweden facing climate change threats and opportunities. Swedish Government Official Reports SOU 2007:60, Stockholm, Sweden, pp. 679. Tardieu, F., 2012: Any trait or trait-related allele can confer drought tolerance: just design the right drought scenario. Journal of Experimental Botany, 63(1), 25-31. Tasker, M.L.(.)., 2008: The Effect of Climate Change on the Distribution and Abundance of Marine Species in the OSPAR Maritime Area. In: ICES Cooperative Research Report No. 293. International Council for the Exploration of the Sea, Copenhagen, Denmark, pp. 1-49. Taylor, S., L. Kumar, N. Reid, and D.J. Kriticos, 2012: Climate change and the potential distribution of an invasive shrub, Lantana camara L. PLoS ONE, 7(4), e35565. Te Linde, A.H., 2007: Effects of climate change on discharge behaviour of the river Rhine[Heinonen, M. (ed.)]. Proceedings of the Third International Conference on Climate and Water, 3 - 6, September, 2007, Helsinki, Finland. Te Linde, A.H., J.C.J.H. Aerts, A.M.R. Bakker, and J.C.J. Kwadijk, 2010a: Simulating low-probability peak discharges for the Rhine basin using resampled climate modeling data. Water Resources Research, 46(3), art. no. W03512. Te Linde, A.H., J.C.J.H. Aerts, and J.C.J. Kwadijk, 2010b: Effectiveness of flood management measures on peak discharges in the Rhine basin under climate change. Journal of Flood Risk Management, 3(4), 248-269. Te Linde, A.H., P. Bubeck, J.E.C. Dekkers, H. De Moel, and J.C.J.H. Aerts, 2011: Future flood risk estimates along the river Rhine. Natural Hazards and Earth System Sciences, 11(2), 459-473. Teich, M., C. Marty, C. Gollut, A. Gret-Regamey, and P. Bebi, 2012: Snow and weather conditions associated with avalanche releases in forests: rare situations with decreasing trends during the last 41years. Cold Regions Science and Technology, 83-84, 77-88. ten Brinke, W.B.M., B. Kolen, A. Dollee, H. van Waveren, and K. Wouters, 2010: Contingency planning for large- scale floods in the Netherlands. Journal of Contingencies and Crisis Management, 18(1), 55-69. ter Hofstede, R., J. Hiddink, and A. Rijnsdorp, 2010: Regional warming changes fish species richness in the Eastern North Atlantic Ocean. Marine Ecology Progress Series, 414, 1-9. Terpstra, T. and J.M. Gutteling, 2008: Households' perceived responsibilities in flood risk management in the Netherlands. International Journal of Water Resources Development, 24(4), 555-565. Tervo, K., 2008: The operational and regional vulnerability of winter tourism to climate variability and change: the case of the Finnish nature-based tourism entrepreneurs. Scandinavian Journal of Hospitality and Tourism, 8(4), 317-332. Thieken, A.H., T. Petrow, H. Kreibich, and B. Merz, 2006: Insurability and mitigation of flood losses in private households in Germany. Risk Analysis, 26(2), 383-395. Thodsen, H., B. Hasholt, and J.H. Kjarsgaard, 2008: The influence of climate change on suspended sediment transport in Danish rivers. Hydrological Processes, 22(6), 764-774. Thodsen, H., 2007: The influence of climate change on stream flow in Danish rivers. Journal of Hydrology, 333(2- 4), 226-238. Thuiller, W., S. Lavergne, C. Roquet, I. Boulangeat, B. Lafourcade, and M.B. Araujo, 2011: Consequences of climate change on the tree of life in Europe. Nature, 470, 531-534. Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Trnka, M., E. Kocmánková, J. Baleka, J. Eitzinger, F. Ruget, H. Formayer, P. Hlavinka, M. Schaumberger, V. Horáková, M. Mo¾ný, and Z. ®aluda, 2010: Simple snow cover model for agrometeorological applications. Agricultural and Forest Meteorology, 150, 1115-1127. Trnka, M., J.E. Olesen, K.C. Kersebaum, A.O. Skjelvag, J. Eitzinger, B. Seguin, P. Peltonen-Sainio, R. Rötter, A. Iglesias, S. Orlandini, M. Dubrovský, P. Hlavinka, J. Balek, H. Eckersten, E. Cloppet, P. Calanca, A. Gobin, V. Vuèetiæ, P. Nejedlik, S. Kumar, B. Lalic, A. Mestre, F. Rossi, J. Kozyra, V. Alexandrov, D. Semerádová, and Z. ®alud, 2011: Agroclimatic conditions in Europe under climate change. Global Change Biology, 17(7), 2298- 2318. Trnka, M., F. Muska, D. Semeradova, M. Dubrovsky, E. Kocmankova, and Z. Zalud, 2007: European Corn Borer life stage model: regional estimates of pest development and spatial distribution under present and future climate. Ecological Modelling, 207(2-4), 61-84. Trnka, M., J. Eitzinger, P. Hlavinka, M. Dubrovska, D. Semerádová, P. Atapanek, S. Thaler, Z. Asalud, M. Molna, and H. Formayer, 2009: Climate-driven changes of production regions in Central Europe. Plant and Soil, 2009(521), 257-266. Troccoli, A., F. Zambon, K. Hodges, and M. Marani, 2012a: Storm surge frequency reduction in Venice under climate change. Climatic Change, 113(3-4), 1065-1079. Troccoli, A., F. Zambon, K.I. Hodges, and M. Marani, 2012b: Reply to comment on "Storm surge frequency reduction in Venice under climate change" by G. Jorda, D. Gomis & M. Marcos. Climatic Change, 113(3-4), 1089-1095. Tsanis, I.K., A.G. Koutroulis, I.N. Daliakopoulos, and D. Jacob, 2011: Severe climate-induced water shortage and extremes in Crete. Climatic Change, 106(4), 667-677. Tu, M., M.J. Hall, P.J.M. de Laat, and M.J.M. de Wit, 2005: Extreme floods in the Meuse river over the past century: aggravated by land-use changes? Physics and Chemistry of the Earth, Parts A/B/C, 30(4-5), 267-276. Tubiello, F.N., J.F. Soussana, and S.M. Howden, 2007: Crop and pasture response to climate change. Proceedings of the National Academy of Sciences of the United States of America, 104(50), 19686-19690. Tuck, G., M.J. Glendining, P. Smith, J.I. House, and M. Wattenbach, 2006: The potential distribution of bioenergy crops in Europe under present and future climate. Biomass and Bioenergy, 30(3), 183-197. Turco, M., M.C. Llasat, J. von Hardenberg, and A. Provenzale, 2013: Impact of climate variability on summer fires in a Mediterranean environment (northeastern Iberian Peninsula). Climatic Change, 116(3-4), 665-678. Uhlmann, B., S. Goyette, and M. Beniston, 2009: Sensitivity analysis of snow patterns in Swiss ski resorts to shifts in temperature, precipitation and humidity under conditions of climate change. International Journal of Climatology, 29(8), 1048-1055. UK National Ecosystem Assessment, 2011: The UK National Ecosystem Assessment: Technical Report. UNEP- WCMC, Cambridge, United Kingdom, pp. 1-1466. UK-ASC, 2011: Adapting to climate change in the UK: measuring progress. Adaptation Sub-Committee, London, United Kingdom. Ulbrich, U., G.C. Leckebusch, and J.G. Pinto, 2009: Extra-tropical cyclones in the present and future climate: a review. Theoretical and Applied Climatology, 96(1-2), 117-131. Ulbrich, U., P. Lionello, D. Belu¹ic, J. Jacobeit, P. Knippertz, F.G. Kuglitsch, G.C. Leckebusch, J. Luterbacher, M. Maugeri, P. Maheras, K.M. Nissen, V. Pavan, J.G. Pinto, H. Saaroni, S. Seubert, A. Toreti, E. Xoplaki, and B. Ziv, 2012: The Climate of the Mediterranean Region. In: Climate of the mediterranean: Synoptic patterns, temperature, precipitation, winds, and their extremes, pp. 301-346. Ulén, B.M. and G.A. Weyhenmeyer, 2007: Adapting regional eutrophication targets for surface waters--influence of the EU Water Framework Directive, national policy and climate change. Environmental Science & Policy, 10(7- 8), 734-742. Unbehaun, W., U. Pröbstl, and W. Haider, 2008: Trends in winter sport tourism: challenges for the future. Tourism Review, 63(1), 36-47. UNEP, 2010: Global Synthesis - A report from the Regional Seas Conventions and Action Plans for the Marine Biodiversity Assessment and Outlook Series. United Nations Environment Programme, Nairobi, Kenya, pp. 1- 55. Usbeck, T., T. Wohlgemuth, M. Dobbertin, C. Pfister, A. Burgi, and M. Rebetez, 2010: Increasing storm damage to forests in Switzerland from 1858 to 2007. Agricultural and Forest Meteorology, 150, 47 55. Van der Linden, P. and J.F.B. Mitchell, 2009: Climate Change and its Impacts: Summary of Research and Results from The ENSEMBLES Project. Met Office Hadley Centre, Exeter, UK, pp. 1-160. Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 van der Velde, M., G. Wriedt, and F. Bouraoui, 2010: Estimating irrigation use and effects on maize yield during the 2003 heatwave in France. Agriculture Ecosystems & Environment, 135(1-2), 90-97. van Dijk, J., N.D. Sargison, F. Kenyon, and P.J. Skuce, 2010: Climate change and infectious disease: helminthological challenges to farmed ruminants in temperate regions. Animal, 4(3), 377-392. Van Nieuwaal, K., P. Driessen, T. Spit, and C. Termeer, 2009: A State of the Art of Governance Literature on Adaptation to Climate Change: Towards A Research Agenda. Report number 003/2009. Knowledge for Climate (KfC), Utrecht, Netherlands, pp. 1-43. van Vliet, M.T.H. and J.J.G. Zwolsman, 2008: Impact of summer droughts on the water quality of the Meuse river. Journal of Hydrology, 353(1-2), 1-17. Van Vliet, M.T.H., J.R. Yearsley, F. Ludwig, S. Vögele, D.P. Lettenmaier, and P. Kabat, 2012: Vulnerability of US and European electricity supply to climate change. Nature Climate Change, 2(9), 676-681. Varakina, Z.L., D.A. Shaposhnikov, B.A. Revich, A.M. Vyazmin, E.D. Yurasova, J. Nurse, and B. Menne, 2011: The impact of air temperature on daily mortality in Archangelsk city in Northwest Russia: a time-series analysis. European Journal of Public Health,. Vautard, R., J. Cattiaux, P. Yiou, J.N. Thepaut, and P. Ciais, 2010: Northern Hemisphere atmospheric stilling partly attributed to an increase in surface roughness. Nature Geoscience, 3(11), 756-761. Veijalainen, N., E. Lotsari, P. Alho, B. Vehviläinen, and J. Käyhkö, 2010: National scale assessment of climate change impacts on flooding in Finland. Journal of Hydrology, 391(3-4), 333-350. Ventrella, D., M. Charfeddine, and M. Bindi, 2012: Agronomic adaptation strategies under climate change for winter durum wheat and tomato in southern Italy: irrigation and nitrogen fertilization. Regional Environmental Change, 12(3), 407-419. Verburg, P.H., D.B. van Berkel, A.M. van Doorn, M. van Eupen, and H.A.R.M. van den Heiligenberg, 2010: Trajectories of land use change in Europe: a model-based exploration of rural futures. Landscape Ecology, 25(2), 217-232. Verny, J. and C. Grigentin, 2009: Container shipping on the Northern Sea Route. International Journal of Production Economics, 122(1), 107-117. Vidal, J.P. and S. Wade, 2009: A multimodel assessment of future climatological droughts in the United Kingdom. International Journal of Climatology, 29(14), 2056-2071. Vilén, T. and P.M. Fernandes, 2011: Forest fires in Mediterranean countries: CO2 emissions and mitigation possibilities through prescribed burning. Environmental Management, 48, 558-567. Villarini, G., J.A. Smith, F. Serinaldi, and A.A. Ntelekos, 2011: Analyses of seasonal and annual maximum daily discharge records for Central Europe. Journal of Hydrology, 399(3-4), 299-312. Vinagre, C., F.D. Santos, H. Cabral, and M.J. Costa, 2011: Impact of climate warming upon the fish assemblages of the Portuguese coast under different scenarios. Regional Environmental Change, 11(4), 779-789. Virkkala, R., R.K. Heikkinen, S. Fronzek, H. Kujala, and N. Leikola, 2013: Does the protected area network preserve bird species of conservation concern in a rapidly changing climate? Biodiversity and Conservation, 22(2), 459-482. Vorogushyn, S., K.E. Lindenschmidt, H. Kreibich, H. Apel, and B. Merz, 2012: Analysis of a detention basin impact on dike failure probabilities and flood risk for a channel-dike-floodplain system along the river Elbe, Germany. Journal of Hydrology, 436-437, 120-131. Vos, C.C., P. Berry, P. Opdam, H. Baveco, B. Nijhof, J. OíHanley, C. Bell, and H. Kuipers, 2008: Adapting landscapes to climate change: examples of climate-proof ecosystem networks and priority adaptation zones. Journal of Applied Ecology, 45(6), 1722-1731. Wade, S.D., J. Rance, and N. Reynard, 2013: The UK climate change risk assessment 2012: assessing the impacts on water resources to inform policy makers. Water Resources Management, 27(4), 1085-1109. Wall, R. and L.S. Ellse, 2011: Climate change and livestock parasites: integrated management of sheep blowfly strike in a warmer environment. Global Change Biology, 17(5), 1770-1777. Walther, G.R., A. Roques, P.E. Hulme, M.T. Sykes, P. Py¹ek, I. Kühn, M. Zobel, S. Bacher, Z. Botta-Dukát, H. Bugmann, B. Czúcz, J. Dauber, T. Hickler, V. Jaro¹ík, M. Kenis, S. Klotz, D. Minchin, M. Moora, W. Nentwig, J. Ott, V.E. Panov, B. Reineking, C. Robinet, V. Semenchenko, W. Solarz, W. Thuiller, M. Vila, K. Vohland, and J. Settele, 2009: Alien species in a warmer world: risks and opportunities. Trends in Ecology & Evolution, 24(12), 686-693. Wamsler, C. and N. Lawson, 2011: The role of formal and informal insurance mechanisms for reducing urban disaster risk: a South-North somparison. Housing Studies, 26(2), 197-223. Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wang, S., R. Mcgrath, T. Semmler, and P. Nolan, 2006: The impact of the climate change on discharge of Suir River Catchment (Ireland) under different climate scenarios. Natural Hazards and Earth System Science, 6(3), 387-395. Wang, S., R. McGrath, J. Hanafin, P. Lynch, T. Semmler, and P. Nolan, 2008: The impact of climate change on storm surges over Irish waters. Ocean Modelling, 25(1-2), 83-94. Ward, D.M., F.M. Cohan, D. Bhaya, J.F. Heidelberg, M. Kuhl, and A. Grossman, 2008: Genomics, environmental genomics and the issue of microbial species. Heredity, 100, 207-219. Ward, P., H. Renssen, J. Aerts, and P. Verburg, 2011: Sensitivity of discharge and flood frequency to twenty-first century and late Holocene changes in climate and land use (River Meuse, northwest Europe). Climatic Change, 106(2), 179-202. Wasowski, J., C. Lamanna, and D. Casarano, 2010: Influence of land-use change and precipitation patterns on landslide activity in the Daunia Apennines, Italy. Quarterly Journal of Engineering Geology and Hydrogeology, 43(4), 387-401. Watkiss, P. and A. Hunt, 2010: Review of Adaptation Costs and Benefit estimates in Europe. [Alistair Hunt and Paul Watkiss Metroeconomica (ed.)]. Report prepared for the European Environment Agency, Bath, UK, pp. 74. Weber, R.W.S., 2009: An evaluation of possible effects of climate change on pathogenic fungi in apple production using fruit rots as examples. Erwerbs-Obstbau, 51(3), 115-120. Wedawatta, G.S.D. and M.J.B. Ingirige, 2012: Resilience and adaptation of small and medium-sized enterprises to flood risk. Disaster Prevention and Management, 21(4), 474-488. Wessel, W., A. Tietema, C. Beier, B. Emmett, J. Penuelas, and T. Riis-Nielson, 2004: A qualitative ecosystem assessment for different shrublands in Western Europe under impact of climate change. Ecosystems, 7, 662-671. Wessolek, G. and S. Asseng, 2006: Trade-off between wheat yield and drainage under current and climate change conditions in northeast Germany. European Journal of Agronomy, 24(4), 333-342. West, J.S., J.A. Townsend, M. Stevens, and B.D.L. Fitt, 2012: Comparative biology of different plant pathogens to estimate effects of climate change on crop diseases in Europe. European Journal of Plant Pathology, 133, 315- 331. Westerhoff, L., E.H. Keskitalo, and S. Juhola, 2011: Capacities across scales: local to national adaptation policy in four European countries. Climate Policy, 11(4), 1071-1085. Wethey, D.S., S.A. Woodin, T.J. Hilbish, S.J. Jones, F.P. Lima, and P.M. Brannock, 2011: Response of intertidal populations to climate: effects of extreme events verses long term change. Journal of Experimental Marine Biology and Ecology, 400(1-2), 132-144. White, M.A., P. Whalen, and G.V. Jones, 2009: Land and wine. Nature Geoscience, 2, 82-84. Whitehead, P.G., R.L. Wilby, R.W. Battarbee, M. Kernan, and A.J. Wade, 2009: A review of the potential impacts of climate change on surface water quality. Hydrological Sciences Journal, 54(1), 101-123. Whittle, R., W. Medd, H. Deeming, E. Kashefi, M. Mort, C. Twigger Ross, G. Walker, and N. Watson, 2010: After the Rain - Learning the Lessons from Flood Recovery in Hull. Final Project Report for 'Flood, Vulnerability and Urban Resilience: a real-time study of local recovery following the floods of June 2007 in Hull'. Lancaster University, Lancaster, United Kingdom, pp. 1-171. WHO, 2008: Heat-health action plans: guidance. World Health Organisation Regional Office for Europe, Copenhagen, Denmark, pp. 1-51. WHO, 2013: Floods in the WHO European region: health effects and their prevention. [Menne, B. and Murray, V. (eds.)]. World Health Organization Regional Office for Europe, Denmark, pp. 1-146. Wiering, M.A. and B.J.M. Arts, 2006: Discursive shifts in Dutch river management: 'deep' institutional change or adaptation strategy? In: Living Rivers: Trends and Challenges in Science and Management . [Leuven, R.S.E.W., A.M.J. Ragas, A.J.M. Smits, and G. Velde (eds.)]. Springer, Netherlands, pp. 327-338. Wilby, R.L., 2008: Constructing climate change scenarios of urban heat island intensity and air quality. Environment and Planning B: Planning and Design, 35(5), 902-919. Wilkinson, P., K.R. Smith, M. Davies, H. Adair, B.G. Armstrong, M. Barrett, N. Bruce, A. Haines, I. Hamilton, T. Oreszczyn, I. Ridley, C. Tonne, and Z. Chalabi, 2009: Public health benefits of strategies to reduce greenhouse- gas emissions: household energy. The Lancet, 374(9705), 1917-1929. Willems, P., J. Olsson, K. Arnbjerg-Nielsen, S. Beecham, A. Pathirana, I. Bülow Gregersen, H. Madsen, and V. Nguyen, 2012: Impacts of Climate Change on Rainfall Extremes and Urban Drainage. IWA Publishing, London, pp. 238. Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wilson, E., 2006: Adapting to climate change at the local level: the spatial planning response. Local Environment, 11(6), 609-625. Wilson, G., 2008: Our knowledge ourselves: Engineers (re)thinking technology in development. Journal of International Development, 20(6), 739-750. Wilson, A.J. and P.S. Mellor, 2009: Bluetongue in Europe: past, present and future. Philosophical Transactions of the Royal Society B-Biological Sciences, 364(1530), 2669-2681. Wilson, D., H. Hisdal, and D. Lawrence, 2010: Has streamflow changed in the Nordic countries? Recent trends and comparisons to hydrological projections. Journal of Hydrology, 394(3 4), 334-346. WLO, 2006: Welvaart en Leefomgeving: Een Scenariostudie voor Nederland in 2040 [in Dutch]. [Janssen, L.H.J.M., Okker, V.R. and Schuur, J. (eds.)]. Central Planning Bureau, Netherlands Environmental Assessment Agency and Spatial Planning Burea, The Hague, Netherlands, pp. 1-239. WMO, 2011: Climate Knowledge for Action: A global Framework for climate Services- Empowering the most vulnerable. WMO-No. 1065. World Meteorological Organization, Geneva. Wong, W.K., B. Stein, E. Torill, H. Ingjerd, and H. Hege, 2011: Climate change effects on spatiotemporal patterns of hydroclimatological summer droughts in Norway. Journal of Hydrometeorology, 12(6), 1205-1220 (doi: 10.1175/2011JHM1357.1). Wreford, A., D. Moran, and N. Adger, 2010: Climate Change and Agriculture. Impacts, Adaptation and Mitigation. OECD, Paris, pp. 135. Yiou, P., P. Ribereau, P. Naveau, M. Nogaj, and R. Brazdil, 2006: Statistical analysis of floods in Bohemia (Czech Republic) since 1825. Hydrological Sciences Journal, 51(5), 930-945. Zachariadis, T., 2010: Forecast of electricity consumption in Cyprus up to the year 2030: The potential impact of climate change. Energy Policy, 38(2), 744-750. Zhou, Q., P.S. Mikkelsen, K. Halsnaes, and K. Arnbjerg-Nielsen, 2012: Framework for economic pluvial flood risk assessment considering climate change effects and adaptation benefits. Journal of Hydrology, 414-415, 539- 549. Zsamboky, M., A. Fernandez-Bilbao, D. Smith, J. Knight, and J. Allan, 2011: Impacts of Climate Change on Disadvantaged UK Coastal Communities. Joseph Rowntree Foundation, York, United Kingdom, pp. 1-62. Zwicke, M., G.A. Alessio, L. Thiery, R. Falcimagne, R. Baumont, N. Rossignol, J. Soussana, and C. Picon Cochard, 2013: Lasting effects of climate disturbance on perennial grassland above ground biomass production under two cutting frequencies. Global Change Biology, in press. Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 23-1: Impacts of climate extremes in the last decade in Europe. Year Region Meteorological Production Systems Agriculture, Health and Environmental Mega- Characteristics and Physical Fisheries, Social Welfare Quality and fire Infrastructure, Forestry, Biological settlements Bioenergy Conservation 2003 Western Hottest summer Damage to road and Grain harvest 35,000 deaths Decline in water Yes and in at least rail transport systems. losses of 20% in August in quality Central 500 years Reduced/ interrupted (Ciais, et al. Central and (Daufresne et al. Europe (Luterbacher et operation of nuclear 2005) Western 2007). al., 2004) power plants (mostly Europe, in France). (Robine et al. High outdoor High transport prices 2008) pollution levels in Rhine due to low (EEA 2012) water levels. 2004/ Iberian Hydrological - Grain harvest - - 2005 Peninsula drought losses of 40% - Portugal (EEA, 2010c) 2007 Southern Hottest summer 1710 buildings burned Approx. Significant Several Yes, - Europe on record in down or 575,500 mortality protected Greece Greece since rendered uninhabitable hectares burnt impact: 6 conservation 1891 (Founda & in Greece (JRC, 2008) area (JRC, deaths in sites (Natura Giannakopoulos 2008) Portugal, 80 2000) were 2009) deaths in destroyed (JRC, Greece. 2008) 2007 England May July Estimated total losses 78 farms 13 deaths and and wettest since 4 billion (3 billion flooded. 48,000 flooded Wales records began in insured losses) Impacts on homes (Pitt, 1766. (Chatterton et al. agriculture 2008). Damage 2010). 50 million costs for health Failure of pumping (Chatterton et effects, incl. station led to 20,000 al. 2010) loss of access to people without water education 287 for 2 weeks million (Chatterton et al. 2010) 2010 Western Hottest summer Fire damage Estimated High outdoor Yes Russia since 1500 to forests 10,000 excess pollution levels (Barriopedro et (Shvidenko et deaths due to in Moscow al., 2011) al., 2011). heatwave in (Bondur, 2011, Reduction in Moscow in July Revich and crop yields and August Shaposhnikov, (Coumou and (Revich and 2012) Rahmstorf, Shaposhnikov, 2012) 2012) 2011 France Hottest and Reduction in snow 8% decline in driest spring in cover for skiing wheat yield France since (AGRESTE, 1880 2011) * Extreme events derived from Coumou and Rahmstorf, 2012. Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 23-2: Selected published cost estimates for planned adaptation in European countries. Region Cost estimate Time period Sectors/Outcomes Reference Europe 2.6-3.5 billion/a In 2100 Coastal adaptation costs Hinkel et al. 2010 Europe 1.7 billion/a By 2020s Protection from river flood risk Rojas et al., in press 3.4 billion/a By 2050s for EU27 7.9 billion/a By 2080s Netherlands 1.2 1.6 billion/a up to 2050 Protection from coastal and Delta Committee, 2008 0.9 1.5 billion/a 2050 2100 river flooding Sweden total of up to 10 2010-2100 Multi sector Swedish Commission billion on Climate and Vulnerability, 2007 Italy 0.4-2 billion By 2080s Coastal protection Bosello et al. 2012, Up to 44 billion Hydrogeological protection Medri et al. 2013. Greece 0.4-3.3 billion Up to 2100 Coastal protection Bank of Greece, 2011 UK 1.8 billion Until 2035 Maintain and improve Thames EA, 2011 2.2 billion 2035-2050 flood protection 7-8 billion At 2100 Renew and improve Thames flood protection New Thames barrier for London Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 23-3: Limits to adaptation to climate change. Area/Location System Adaptation Limits to adaptation measure(s) References measures Low altitude/ small- Ski tourism Artificial Climatic, technological and Landauer et al, 2012 ; size ski resorts snowmaking environmental constraints Steiger, 2010; Steiger, Economic viability 2011; Steiger and Social acceptability of charging for Mayer, 2008, Unbehaun previously free skiing. et al., 2008 Social acceptability of alternatives for winter sport/leisure. Thermal power Once-through Closed- circuit High investment cost for retrofitting van Vliet et al., 2012, plants/ cooling cooling cooling existing plants Koch and Vögele, 2009, through river intake systems Hoffman et al., 2013 and discharge Reduced load Increased transport prices (Rhine and Jonkeren, 2009, factor of inland Moselle market) Jonkeren et al., 2007 Rivers used for Inland ships freight transport transport Use of smaller Existing barges below optimal size Demirel, 2011 ships (Rhine) Agriculture, Northern Arable crops Sowing date as Other constraints (e.g. frost) limit Oort, 2012 and Continental agricultural farmer behaviour Europe. adaptation Agriculture, Northern Arable crops Irrigation Groundwater availability, competition Olesen et al., 2011 and Continental with other users. Europe. Agriculture, High value Change Legislation on cultivar and Box 23-1 Viticulture crops distribution geographical region Conservation Alpine Extend habitat No technological adaptation option. Engler et al., 2011, Cultural landscapes meadow/ Dullinger et al., 2012 Conservation of Movement of Extend habitat Landscape barriers and absence of Butchart et al., 2010, species richness species climate projections in selection of Araújo et al., 2011; Filz conservation areas. et al., 2012; Virkkala et al., 2013 Forests Movement of Introduce new Not socially acceptable, Giuggiola et al., 2010; species and species Legal barriers to non-native species Hemery et al., 2010; productivity García-López J.M. and reduction Alluéa, 2011, Casalegno et al., 2007 Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 23-4: Assessment of climate change impacts by sub-region by 2050, assuming a medium emissions scenario, and no planned adaptation. Impacts assume economic development, including land use change. Impacts are assessed for the whole sub-region, although differences in impact within sub-regions are estimated for some impacts. Alpine Southern Northern Continental Atlantic Energy 1 Wind energy 23.3.4 production 2 Hydropower 23.3.4 generation Thermal power 23.3.4, 8.2.3.2 production Energy consumption 23.3.4, 23.8.1 (net annual change) Transport Road accidents 3 23.3.3 Rail delays (weather- ? ? ? 23.3.3, 8.3.3.6 4 related) Load factor of inland ? ? ? 23.3.3 ships Transport time and ? ? ? 23.3.3, 18.3.3.3.5 cost in ocean routes Settlements River flood damages 23.3.1 Coastal flood n/a 23.3.1 damages Tourism Length of ski season ? ? 23.3.6, 3.5.7 Human health Heat wave mortality 23.5.1 and morbidity Food safety 23.5.1 Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Social and cultural Impacts Social costs of floods 23.5.3 Damage on cultural buildings 23.5.4 Loss of cultural ? 23.5.4 landscapes Environmental quality Air quality (ozone ? 23.6.1 background levels) Air quality ? 23.6.1 (particulates) Water quality 23.6.3 Key: Green means a beneficial change Red means a harmful change ? means no relevant literature found Confidence levels: Risks were identified based on assessment of the literature and expert judgment. Footnotes: 1 Simulations have been performed, but mostly for the period after 2070. 2 The increasing trend is for Norway. 3 The decreasing trend refers mainly to the number of severe accidents. 4 Impacts have been studied and quantified for UK only. The increasing trend stands for summer delays and the decreasing trend for winter delays. 5 In both seasons, no significant impacts are expected by 2020, while more substantial changes are expected by 2080. For 2050 impacts are assumed to vary linearly (although this may not be the case). 6 The constant trend stands for the Mediterranean, where some studies estimate no changes due to climate change at least until 2030 or even 2060. Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 23-5: Key risks from climate change in Europe and the potential for reducing risk through mitigation and adaptation. Risk levels are presented in three timeframes: the present, near-term (2030-2040), and longer-term (2080-2100). For each timeframe, risk levels are estimated for a continuation of current adaptation and for a hypothetical highly adapted state. For a given key risk, change in risk level through time and across magnitudes of climate change is illustrated, but because the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks, sectors, or regions. Key risks were identified based on assessment of the literature and expert judgment. Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 23-6: Observed changes in key indicators in ecological and human systems attributable to climate factors. Confidence in Confidence in attribution to change Indicator Change in indicator Key references Section detection in climate factors [**] Bio-Physical Systems Glacier Fast mass loss of 30 Swiss High confidence Medium confidence Huss, 2010 18.3.1.3 retreat glaciers since the 1980s WG1 10.5 Infrastructure Storm Increase since 1970s High confidence No causal role for Barredo, 2010 23.3.7 losses climate Hail Increase in parts of Germany Low confidence Low confidence Kunz et al., 2009 23.3.7 losses Flood Increasing general trend in Medium confidence No causal role for Barredo, 2009; 23.3.1 losses economic losses in Europe climate Barredo et al., 2012 since 1970s; none in some locations Agriculture, Fisheries, Forestry, and Bioenergy Production C3 crop CO2 induced positive High confidence High confidence Amthor, 2001; 7.2.1 yield contribution to yield since (high agreement, (high agreement, Long et al., 2006; preindustrial for C3 crops robust evidence) robust evidence) McGrath and Lobell, 2011 Wheat Stagnation of wheat yields in High confidence Medium confidence Lobell et al. 2011 ; 23.4.1 yield some countries in recent Brisson et al., 2010; decades Kristensen et al., 2011 Phenolog Earlier greening, Earlier leaf High confidence High confidence Menzel et al., 2006 4.4.1.1 y leaf emergence and fruit set in (high agreement, (high agreement, greening temperate and boreal climate, robust evidence) robust evidence) Phytopla Increased phytoplankton High confidence Medium confidence Beaugrand et al., 6.3.2 nkton productivity in NE. Atlantic, 2002; productiv decrease in warmer regions, due Edwards and ity to warming trend and Richardson, 2004 hydroclimatic variations Ocean Northward movement of High confidence Medium confidence Philippart et al. , 2011 6.3.2 systems species and increased species richness due to warming trend Environmental quality and biodiversity Biodivers Increased number of Medium confidence Medium confidence Walther et al., 2009 4.2.4.6 ity colonization events by alien (high agreement, plant species in Europe medium evidence) Migrator Decline over the period 1990- Medium confidence Medium confidence Moller et al., 2008 4.4.1.1 y birds 200 of species that did not (medium agreement, advance their spring migration medium evidence) Tree Upward shift in tree line in Medium evidence Medium confidence Gehrig-Fasel et al., 18.3.2.1 species Europe (medium agreement, 2007, Lenoir et al., high evidence) 2008 Forest Increase in burnt area High confidence High confidence Camia and Amatulli fires (high agreement, 2009; Hoinka et al., 23.4.4 robust evidence) 2009; Carvalho et al., 2010; Salis et al., in press; Pereira et al., 2005; Koutsias et al., 2012 NOTE: The studies included in this table are those with good evidence of a detection of a long term trend in the outcome of interested, and where there has been an assessment of the attribution of the trend to an observed change in climate factor. It is not possible to make an attribution to anthropogenic climate change at this scale see chapter 18 for a more complete discussion. Subject to Final Copyedit 85 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 BOX 23-1 TABLE Alpine Atlantic Continental Northern Southern Provisioning services: No (1) (1) Food production (1) (1) (1) (4) (1) No (1) Livestock production (1) Fibre production (1) (1) Bioenergy production (1) (1) No (1) No (1) No (1) Fish production (1) (1) (1) (2) (5) (1) (2) (6) Timber production No (2) No (2) V (2) No (3) No (1) (5) (1) (1) Non-wood forest products (1) No (1) (6) (2) (1) (9) Sum of effects on No (1) No (4) No (4) No (2) No (3) provisioning services (7) (11) (2) (3) (2) Regulating services: Climate regulation (carbon sequestration) (4) (4) (4) (3) (3) - General/forests No (1) No (1) No (1) No (1) (1) (3) (1) No (1) No (1) No (1) - Wetland (1) (1) (1) (1) No (1) No (1) No (1) No (1) - Soil carbon stocks (3) (2) (2) (1) (1) Pest control (1) (1) (1) (1) Natural hazard regulation* - Forest fires / wildfires (1) (2) (1) - Erosion, avalanche, (2) landslide (1) - Flooding (1) No (1) - Drought (1) (1) Water quality regulation (1) (1) (2) (2) (2) (3) (1) Biodiversity No (1) (4) (4) (2) (8) (4) (9) (6) (6) (8) (4) Sum of effects on regulating No (2) No (4) No (2) No (2) No (3) services (11) (9) (9) (8) (14) Cultural services: Recreation (fishing, nature (1) (1) (1) enjoyment) (2) Tourism (skiing) (1) (1) Subject to Final Copyedit 86 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Alpine Atlantic Continental Northern Southern Asthetic/heritage (landscape No (1) character, cultural (1) (1) (1) (1) landscapes) Sum of effects on cultural (1) (1) No (1) (1) (2) services (1) (1) (1) (3) Key: = decreasing impacts = increasing impacts No = neutral effect (1) = number in brackets refers to the number of studies supporting the change (increasing, decreasing, neutral) in ecosystem service. Footnotes: * A decline in ecosystem services implies an increased risk of the specified natural hazard ^ Entries for biodiversity are those that were found during the literature search for climate change impacts on ecosystem services. A wider discussion of the impacts of climate change on biodiversity can be found in Section 23.6 and AR5 4.3.4. References: Albertson et al., 2010; Bastian, 2013; Bolte et al., 2009; Briner et al., 2012; Canu et al., 2010; Civantos et al., 2012; Clark et al., 2010a; Forsius et al., 2013; Fuhrer et al., 2006; Garcia-Fayos and Bochet, 2009; Gret-Regamy et al., 2008; Gret-Regamy et al., 2013; Hemery, 2008; Johnson et al., 2009; Koca et al., 2006; Lindner et al., 2010; Lorz et al., 2010; Metzger et al., 2008; Milad et al., 2011; Okruszko et al., 2011; Palahi et al., 2008; Rusch, 2012; Schroter et al., 2005; Seidl et al., 2011; Seidl and Lexer, 2013; Wessel et al., 2004. Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 23-1: Sub-regional classification of the IPCC Europe region. Based on Metzger et al., 2005. Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 89 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 23-2: First row: Projected changes in the mean number of heat waves occurring in the months May to September for the period 2071-2100 compared to 1971-2000 (number per 30 years). Heat waves are defined as periods of more than 5 consecutive days with daily maximum temperature exceeding the mean maximum temperature of the May to September season of the control period (1971-2000) by at least 5°C. Second and third rows: Projected seasonal changes in heavy precipitation defined as the 95th percentile of daily precipitation (only days with precipitation > 1mm/day are considered) for the period 2071-2100 compared to 1971-2000 (in %) in the months December to January (DJF) and June to August (JJA). Fourth row: Projected changes in the 95th percentile of the length of dry spells for the period 2071-2100 compared to 1971-2000 (in days). Dry spells are defined as periods of at least 5 consecutive days with daily precipitation below 1mm. Hatched areas indicate regions with robust (at least 66% of models agree in the sign of change) and/or statistical significant change (significant on a 95% confidence level using Mann-Whitney-U test). For the eastern parts of Black Sea, Eastern Anatolia and Southeast Anatolia (Turkey), no regional climate model projections are available. Changes represent the mean over 8 (RCP4.5, left side) and 9 (RCP8.5, right side) regional model simulations compiled within the EURO-CORDEX initiative. Adapted from Jacob et al. (2013). [Illustration to be redrawn to conform to IPCC publication specifications.] !"#$%&'()*$+($*,*-./0-0.1$2*3'(2$24*$.+$ -,03'.*$-&'()*$ !"#$%&'()*(+(,-./.$0()*1(2,!3(4'/'5%"(&5'/6( !"#$%&'()*(+(,-./.$0()*1(-./7'54'/-'(8#&9(2,!3:*;<;( !"#$%&'()*(+(,-./.$0()*1(-./7'54'/-'(8#&9(2,!3:*=;;( !"#$%&'(>*(+(,-./.$0(>*1(2,!3(4'/'5%"(&5'/6( !"#$%&'(>*(+(,-./.$0(>*1(-./7'54'/-'(8#&9(2,!3:*;<;( Figure 23-3: Percentage change in electricity demand in Greece attributable to climate change, under a range of climate scenarios and economic assumptions. Source: Mirasgedis et al., 2007. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 90 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 23-4: Percentage change in simulated water-limited yield for winter wheat in 2030 with respect to the 2000 baseline for the A1B scenario using ECHAM5 (left column) and HadCM3 (right) GCMs. Upper maps to do not take adaptation into account. Bottom maps include adaptation. Source: Donatelli et al., 2012. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 91 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 23-5: Changes in forest fire risk in Europe for two time periods: baseline (left) and 2041 2070 (right), based on high-resolution regional climate models and the SRES A1B emission scenario. Source: Lung et al., 2013. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 92 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 23 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 23-6: Adaptation and mitigation options and their effects on biodiversity. The horizontal axis ranges from positive effects on biodiversity (left-hand side) to negative effects (right-hand side). Each mitigation/adaptation option is located on the biodiversity effect axis (solid bars), including an estimate of the uncertainties associated with the assessment (error bars). The various options are given vertically with mitigation at the top and adaptation at the bottom. Options located toward the centre of the vertical axis have benefits for both mitigation and adaptation. Thus, options located at the centre left of the figure have benefits for mitigation, adaptation and biodiversity and hence are labelled as win-win-win . Other combinations of benefits and dis-benefits are labelled accordingly, e.g. win-lose-win, lose-win-lose, etc. Based on Paterson et al., 2009. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 93 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 24. Asia Coordinating Lead Authors Yasuaki Hijioka (Japan), Erda Lin (China), Joy Jacqueline Pereira (Malaysia) Lead Authors Richard T. Corlett (China), Xuefeng Cui (China), Gregory Insarov (Russian Federation), Rodel Lasco (Philippines), Elisabet Lindgren (Sweden), Akhilesh Surjan (India) Contributing Authors Elena M. Aizen (USA), Vladimir B. Aizen (USA), Rawshan Ara Begum (Bangladesh), Kenshi Baba (Japan), Monalisa Chatterjee (USA / India), J. Graham Cogley (Canada), Noah Diffenbaugh (USA), Li Ding (Singapore), Qingxian Gao (China), Matthias Garschagen (Germany), Masahiro Hashizume (Japan), Manmohan Kapshe (India), Andrey G. Kostianoy (Russia), Kathleen McInnes (Australia), Sreeja Nair (India), S.V.R.K. Prabhakar (India), Yoshiki Saito (Japan), Andreas Schaffer (Singapore), Rajib Shaw (Japan), Dáithí Stone (Canada / South Africa / USA), Reiner Wassman (Philippines), Thomas J. Wilbanks (USA), Shaohong Wu (China) Review Editors Rosa Perez (Philippines), Kazuhiko Takeuchi (Japan) Volunteer Chapter Scientist Yuko Onishi (Japan), Wen Wang (China) Contents Executive Summary 24.1. Introduction 24.2. Major Conclusions from Previous Assessments 24.3. Observed and Projected Climate Change 24.3.1. Observed Climate Change 24.3.2. Projected Climate Change 24.4. Observed and Projected Impacts, Vulnerabilities, and Adaptation 24.4.1. Freshwater Resources 24.4.1.1. Sub-Regional Diversity 24.4.1.2. Observed Impacts 24.4.1.3. Projected Impacts 24.4.1.4. Vulnerabilities to Key Drivers 24.4.1.5. Adaptation Options 24.4.2. Terrestrial and Inland Water Systems 24.4.2.1. Sub-Regional Diversity 24.4.2.2. Observed Impacts 24.4.2.3. Projected Impacts 24.4.2.4. Vulnerabilities to Key Drivers 24.4.2.5. Adaptation Options 24.4.3. Coastal Systems and Low-Lying Areas 24.4.3.1. Sub-Regional Diversity 24.4.3.2. Observed Impacts 24.4.3.3. Projected Impacts 24.4.3.4. Vulnerabilities to Key Drivers Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 24.4.3.5. Adaptation Options 24.4.4. Food Production Systems and Food Security 24.4.4.1. Sub-Regional Diversity 24.4.4.2. Observed Impacts 24.4.4.3. Projected Impacts 24.4.4.4. Vulnerabilities to Key Drivers 24.4.4.5. Adaptation Options 24.4.5. Human Settlements, Industry, and Infrastructure 24.4.5.1. Sub-Regional Diversity 24.4.5.2. Observed Impacts 24.4.5.3. Projected Impacts 24.4.5.4. Vulnerabilities to Key Drivers 24.4.5.5. Adaptation Options 24.4.6. Human Health, Security, Livelihoods, and Poverty 24.4.6.1. Sub-Regional Diversity 24.4.6.2. Observed Impacts 24.4.6.3. Projected Impacts 24.4.6.4. Vulnerabilities to Key Drivers 24.4.6.5. Adaptation Options 24.4.7. Valuation of Impacts and Adaptation 24.5. Adaptation and Managing Risks 24.5.1. Conservation of Natural Resources 24.5.2. Flood Risks and Coastal Inundation 24.5.3. Economic Growth and Equitable Development 24.5.4. Mainstreaming and Institutional Barriers 24.5.5. Role of Higher Education in Adaptation and Risk Management 24.6. Adaptation and Mitigation Interactions 24.7. Intra-regional and Inter-regional Issues 24.7.1. Trans-boundary Pollution 24.7.2. Trade and Economy 24.7.3. Migration and Population Displacement 24.8. Research and Data Gaps 24.9. Case Studies 24.9.1. Transboundary Adaptation Planning and Management - Lower Mekong River Basin 24.9.2. Glaciers of Central Asia References Chapter Box 24-1. What's New on Asia in AR5? Frequently Asked Questions 24.1: What will the projected impact of future climate change be on freshwater resources in Asia? 24.2: How will climate change affect food production and food security in Asia? 24.3: Who is most at risk from climate change in Asia? Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Executive Summary Warming trends and increasing temperature extremes have been observed across most of the Asian region over the past century (high confidence) [24.3]. Increasing numbers of warm days and decreasing numbers of cold days have been observed, with the warming trend continuing into the new millennium. Precipitation trends including extremes are characterized by strong variability, with both increasing and decreasing trends observed in different parts and seasons of Asia. Water scarcity is expected to be a major challenge for most of the region due to increased water demand and lack of good management (medium confidence) [24.4.3]. Water resources are important in Asia because of the massive population and vary among regions and seasons. However, there is low confidence in future precipitation projections at a subregional scale and thus in future freshwater availability in most parts of Asia. Population growth and increasing demand arising from higher standards of living could worsen water security in many parts in Asia and affect many people in future. Integrated water management strategies could help adapt to climate change, including developing water saving technologies, increasing water productivity, and water reuse. The impacts of climate change on food production and food security in Asia will vary by region with many regions to experience a decline in productivity (medium confidence) [24.4.4]. This is evident in the case of rice production. Most models, using a range of GCMs and SRES scenarios, show that higher temperatures will lead to lower rice yields as a result of shorter growing periods. There are a number of regions that are already near the heat stress limits for rice. However, CO2 fertilization may at least in part offset yield losses in rice and other crops. In Central Asia, some areas could be winners (cereal production in northern and eastern Kazakhstan could benefit from the longer growing season, warmer winters and slight increase in winter precipitation), while others could be losers (western Turkmenistan and Uzbekistan, where frequent droughts could negatively affect cotton production, increase water demand for irrigation, and exacerbate desertification). In the Indo-Gangetic Plains of South Asia there could be a decrease of about 50% in the most favorable and high yielding wheat area due to heat stress at 2x CO2. Sea- level rise will inundate low lying areas and will especially affect rice growing regions. There are many potential adaptation strategies being practiced and being proposed but research studies on their effectiveness are still few. Terrestrial systems in many parts of Asia have responded to recent climate change with shifts in the phenologies, growth rates, and the distributions of plant species, and permafrost degradation, and the projected changes in climate during the 21st Century will increase these impacts (high confidence) [24.4.2]. Boreal trees will likely invade treeless arctic vegetation, while evergreen conifers will likely invade deciduous larch forest. Large changes may also occur in arid and semiarid areas, but uncertainties in precipitation projections make these more difficult to predict. The rates of vegetation change in the more densely populated parts of Asia may be reduced by the impact of habitat fragmentation on seed dispersal, while the impacts of projected climate changes on the vegetation of the lowland tropics are currently poorly understood. Changes in animal distributions have also been projected, in response to both direct impacts of climate change and indirect impacts though changes in the availability of suitable habitats. Coastal and marine systems in Asia are under increasing stress from both climatic and non-climatic drivers (high confidence) [24.4.3]. It is likely that mean sea-level rise will contribute to upward trends in extreme coastal high water levels [WG1 Section 3.7.6]. In the Asian Arctic, rising sea-levels are expected to interact with projected changes in permafrost and the length of the ice-free season to cause increased rates of coastal erosion (high agreement, medium evidence). Mangroves, salt marshes and seagrass beds may decline unless they can move inland, while coastal freshwater swamps and marshes will be vulnerable to saltwater intrusion with rising sea-levels. Widespread damage to coral reefs correlated with episodes of high sea-surface temperature has been reported in recent decades and there is high confidence that damage to reefs will increase during the 21st century as a result of both warming and ocean acidification. Marine biodiversity is expected to increase at temperate latitudes as warm- water species expand their ranges northwards (high confidence), but may decrease in the tropics if thermal tolerance limits are exceeded (medium confidence). Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Multiple stresses caused by rapid urbanization, industrialization and economic development will be compounded by climate change (high confidence) [24.4, 24.5, 24.6, 24.7]. Climate change is expected to adversely affect the sustainable development capabilities of most Asian developing countries by aggravating pressures on natural resources and the environment. Development of sustainable cities in Asia with fewer fossil fuel driven vehicles and with more trees and greenery would have a number of co-benefits, including improved public health. Extreme climate events will have an increasing impact on human health, security, livelihoods, and poverty, with the type and magnitude of impact varying across Asia (high confidence) [24.4.6]. More frequent and intense heat-waves in Asia will increase mortality and morbidity in vulnerable groups. Increases in heavy rain and temperature will increase the risk of diarrheal diseases, dengue fever and malaria. Increases in floods and droughts will exacerbate rural poverty in parts of Asia due to negative impacts on the rice crop and resulting increases in food prices and the cost of living. Studies of observed climate changes and their impacts are still inadequate for many areas, particularly in North, Central and West Asia (high confidence) [24.8]. Improved projections for precipitation, and thus water supply, are most urgently needed. Understanding of climate change impacts on ecosystems in Asia is currently limited by the incompleteness and inaccessibility of biodiversity information. Major research gaps in the tropics include the temperature dependence of carbon fixation by tropical trees and the thermal tolerances and acclimation capacities of both plants and animals. Interactions between climate change and the direct impacts of rising CO2 on crops and natural ecosystems are also currently poorly understood. More research is needed on impacts, vulnerability and adaptation in urban settlements, especially cities with populations under 500,000. More generally, there is a need to develop low-cost adaptation measures appropriate to the least developed parts of the region. 24.1. Introduction Asia is defined here as the land and territories of 51 countries/regions (see Figure 24-1). It can be broadly divided into six subregions based on geographical position and coastal peripheries. These are (in alphabetical order) Central Asia (5 countries), East Asia (7 countries/regions), North Asia (2 countries), South Asia (8 countries), Southeast Asia (12 countries) and West Asia (17 countries). The population of Asia was reported to be about 4,299 million in 2013, which is about 60% of the world population (UN, 2013). The population density was reportedly about 134 per square kilometer in 2012 (PRB, 2012). The highest life expectancy at birth is 84 (Japan) and the lowest is 50 (Afghanistan) (CIA, 2013). The GDP per capita ranged from US$620 (Afghanistan for 2011) to US$46,720 (Japan for 2012) (World Bank, 2013). [INSERT FIGURE 24-1 HERE Figure 24-1: The land and territories of 51 countries in Asia. NOTE: Currently in production and will be brought to specification using the current UN-accepted maps.] 24.2. Major Conclusions from Previous Assessments Major highlights from previous assessments for Asia include: Warming trends including higher extremes are strongest over the continental interiors of Asia, and warming in the period 1979 onwards was strongest over China in winter, and northern and eastern Asia in spring and autumn (see WGI AR4 Section 3.2.2.7 and SREX Section 3.3.1). From 1900 to 2005, precipitation increased significantly in northern and central Asia but declined in parts of southern Asia (see WGI AR4 SPM). Future climate change is likely to affect water resource scarcity with enhanced climate variability and more rapid melting of glaciers (see WGII AR4 Section 10.4.2) Increased risk of extinction for many plant and animal species in Asia is likely as a result of the synergistic effects of climate change and habitat fragmentation (see WGII AR4 Section 10.4.4). Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Projected sea-level rise is very likely to result in significant losses of coastal ecosystems (see WGII AR4 Sections 10.4.3.2, 10.6.1). There will be regional differences within Asia in the impacts of climate change on food production (see WGII AR4 Section 10.4.1.1). Due to projected sea-level rise, a million or so people along the coasts of South and Southeast Asia will likely be at risk from flooding (high confidence, see WGII AR4 Section 10.4.3.1). It is likely that climate change will impinge on sustainable development of most developing countries of Asia as it compounds the pressures on natural resources and the environment associated with rapid urbanisation, industrialisation and economic development (see WGII AR4 Section 10.7). Vulnerabilities of industry, infrastructure, settlements and society to climate change are generally greater in certain high-risk locations, particularly coastal and riverine areas (see WGII AR4 Section 7.3, 7.4, 7.5). _____ START BOX 24-1 HERE _____ Box 24-1. What s New on Asia in AR5? Improved country coverage on observed and future impacts of climate change Increase in the number of studies reflecting advances in research tools (e.g. more use of remote sensing and modeling of impacts); with an evaluation of detection and attribution where feasible. More conclusions have confidence statements, while confidence levels have changed in both directions since AR4. Expanded coverage of issues: for example discussion of the Himalayas has been expanded to cover observed and projected impacts (see Box 3-2), including those on: tourism (see Section10.6.2); livelihood assets such as water and food (see Sections 9.3.3.1, 13.3.1.1, 18.5.3 and 19.6.3); poverty (see Section 13.3.2.3); culture (see Sections 12.3.2); flood risks (see Sections 18.3.1.1 and 24.2.1); health risks (see Section 24.4.6.2); and ecosystems (see Section 24.4.2.2). _____ END BOX 24-1 HERE _____ 24.3. Observed and Projected Climate Change 24.3.1. Observed Climate Change Temperature. It is very likely that mean annual temperature has increased over the past century over most of the Asia region, but there are areas of the interior and at high latitudes where the monitoring coverage is insufficient for the assessment of trends (see WGI AR5 Chapter 2, Figure 24-2). New analyses continue to support the AR4 and SREX conclusions that it is likely that the numbers of cold days and nights have decreased and the numbers of warm days and nights have increased across most of Asia since about 1950, and heat wave frequency has increased since the middle of the 20th century in large parts of Asia (see WGI AR5 Section 2.6.1). As a part of the polar amplification, large warming trends (>2°C per 50 years) in the second half of the 20th century were observed in the northern Asian sector (see WGI AR5 Section 14.8.8). Over the period 1901-2009, the warming trend was particularly strong in the cold season between November and March, with an increase of 2.4°C in the mid- latitude semi-arid area of Asia (see WGI AR5 Section 14.8.8). Increasing annual mean temperature trends at the country scale in East and South Asia have been observed during the 20th century (Table 24-SM-1). In West Asia, upward temperature trends are notable and robust in recent decades (see WGI AR5 Section 14.8.10). Across Southeast Asia, temperature has been increasing at a rate of 0.14°C to 0.20°C per decade since the 1960s, coupled with a rising number of hot days and warm nights, and a decline in cooler weather (see WGI AR5 Section 14.8.12). Precipitation and Monsoons. Most areas of the Asian region lack sufficient observational records to draw conclusions about trends in annual precipitation over the past century (see WGI AR5 Chapter 2, Figure 24-2, Table 24-SM-2). Precipitation trends, including extremes, are characterized by strong variability, with both increasing and decreasing trends observed in different parts and seasons of Asia (see WGI AR5 Chapter 14 and Table 24-SM-2). In Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 northern Asia, the observations indicate some increasing trends of heavy precipitation events, but in central Asia, no spatially coherent trends were found (see WGI AR5 Section 14.8.8). Both the East Asian summer and winter monsoon circulations have experienced an interdecadal scale weakening after the 1970s, due to natural variability of the coupled climate system, leading to enhanced mean and extreme precipitation along the Yangtze River valley (30°N), but deficient mean precipitation in North China in summer (see WGI AR5 Section 14.8.9). A weakening of the East Asian summer monsoon since the 1920s was also found in sea level pressure gradients (low confidence, see WGI AR5 Section 2.7.4). In West Asia, a weak but non-significant downward trend in mean precipitation was observed in recent decades, although with an increase in intense weather events (see WGI AR5 Section 14.8.10). In South Asia, seasonal mean rainfall shows interdecadal variability, noticeably a declining trend with more frequent deficit monsoons under regional inhomogeneities (see WGI AR5 Section 14.8.11). Over India, the increase in the number of monsoon break days and the decline in the number of monsoon depressions are consistent with the overall decrease in seasonal mean rainfall (see WGI AR5 Section 14.8.11). But an increase in extreme rainfall events occurred at the expense of weaker rainfall events over the central Indian region and in many other areas (see WGI AR5 Section 14.2.2.1). In South Asia, the frequency of heavy precipitation events is increasing, while light rain events are decreasing (see WGI AR5 Section 14.8.11). In Southeast Asia, annual total wet-day rainfall has increased by 22 mm per decade, while rainfall from extreme rain days has increased by 10 mm per decade, but climate variability and trends differ vastly across the region and between seasons (see WGI AR5 Section 14.4.12, 14.8.12). In Southeast Asia, between 1955 and 2005 the ratio of rainfall in the wet to the dry seasons increased. While an increasing frequency of extreme events has been reported in the northern parts of Southeast Asia, decreasing trends in such events are reported in Myanmar (see WGI AR5 Section 14.4.12). In Peninsular Malaya during the southwest monsoon season, total rainfall and the frequency of wet days decreased, but rainfall intensity increased in much of the region. On the other hand, during the northeast monsoon, total rainfall, the frequency of extreme rainfall events, and rainfall intensity all increased over the peninsula (see WGI AR5 Section 14.4.12). Tropical and Extratropical Cyclones. Significant trends in tropical cyclones making landfall are not found on shorter timescales. Time series of cyclone indices show weak upward trends in the western North Pacific since the late 1970s, but interpretation of longer-term trends is constrained by data quality concerns (see WGI AR5 Section 2.6.3). A decrease in extratropical cyclone activity and intensity over the last 50 years has been reported for northern Eurasia (60°N-40°N), including lower latitudes in East Asia (see WGI AR5 Section 2.6.4). Surface Wind Speeds. Over land in China, including the Tibetan region, a weakening of the seasonal and annual mean winds, as well as the maximums, is reported from around the 1960s or 1970s to the early 2000s (low confidence, see WGI AR5 Section 2.7.2). Oceans. A warming maximum is observed at 25°N-65°N with signals extending to 700 m depth and is consistent with poleward displacement of the mean temperature field (see WGI AR5 Section 3.2.2). The pH measurements between 1983 and 2008 in the western North Pacific showed a -0.0018 +/- 0.0002 yr-1 decline in winter and -0.0013 +/- 0.0005 yr-1 decline in summer (see WGI AR5 Section 3.8.2). Over the period 1993-2010, large rates of sea-level rise in the western tropical Pacific were reported, corresponding to an increase in the strength of the trade winds in the central and eastern tropical Pacific (see WGI AR5 Section 13.6.1). Spatial variation in trends in Asian regional sea level may also be specific to a particular sea or ocean basin. For example, a rise of 5.4 +/- 0.3 mm yr 1 in the Sea of Japan from 1993 to 2001 is nearly two times the GMSL trend, with more than 80% of this rise being thermosteric, and regional changes of sea level in the Indian Ocean that have emerged since the 1960s are driven by changing surface winds associated with a combined enhancement of Hadley and Walker cells (see WGI AR5 Section 13.6.1). 24.3.2. Projected Climate Change The AR4 assessed that warming is very likely in the 21st century (Christensen et al., 2007), and that assessment still holds for all land areas of Asia in the mid- and late-21st-century, based on the CMIP5 simulations under all four RCP scenarios (Figures 24-2, 24-SM-1, and Table 24-SM-3). Ensemble-mean changes in mean annual temperature exceed 2C above the late-20th-century baseline over most land areas in the mid-21st-century under RCP8.5, and range from greater than 3C over South and Southeast Asia to greater than 6C over high latitudes in the late-21st- Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 century. The ensemble-mean changes are less than 2C above the late-20th-century baseline in both the mid- and late-21st-century under RCP2.6, with the exception of changes between 2C and 3C over the highest latitudes. Projections of future annual precipitation change are qualitatively similar to those assessed in the AR4 (Christensen et al., 2007) (Figure 24-2). Precipitation increases are very likely at higher latitudes by the mid-21st-century under the RCP8.5 scenario, and over eastern and southern areas by the late-21st-century. Under the RCP2.6 scenario, increases are likely at high latitudes by the mid-21st century, while it is likely that changes at low latitudes will not substantially exceed natural variability. [INSERT FIGURE 24-2 HERE Figure 24-2: Observed and projected changes in annual average temperature and precipitation in Asia.] Tropical and Extra-Tropical Cyclones. The future influence of climate change on tropical cyclones is likely to vary by region, but there is low confidence in region-specific projections of frequency and intensity. However, better process understanding and model agreement in specific regions indicate that precipitation will likely be more extreme near the centers of tropical cyclones making landfall in West, East, South and Southeast Asia (see WGI AR5 Sections 14.6, 14.8.9, 14.8.10, 14.8.11, 14.8.12). There is medium confidence that a projected poleward shift in the North Pacific storm track of extratropical cyclones is more likely than not. There is low confidence in the magnitude of regional storm track changes and the impact of such changes on regional surface climate (see WGI AR5 Section 14.6) Monsoons. Future increases in precipitation extremes related to the monsoon are very likely in East, South and Southeast Asia (see WGI AR5 Sections 14.2.1, 14.8.9, 14.8.11, 14.8.12). More than 85% of CMIP5 models show an increase in mean precipitation in the East Asian summer monsoons, while more than 95% of models project an increase in heavy precipitation events (see WGI AR5 Section 14.2.2 and Figure 14.4). All models and all scenarios project an increase in both the mean and extreme precipitation in the Indian summer monsoon (see WGI AR5 Section 14.2.2 and SAS in Figure 14.4). In these two regions, the interannual standard deviation of seasonal mean precipitation also increases (see WGI AR5 Section 14.2.2). Oceans. The ocean in subtropical and tropical regions will warm in all RCP scenarios and will show the strongest warming signal at the surface (see WGI AR5 Section 12.4.7 and Figure 12.12). Negligible change or a decrease in mean significant wave heights are projected for the trade and monsoon wind regions of the Indian Ocean (see WGI AR5 Section 13.7.3). 24.4. Observed and Projected Impacts, Vulnerabilities, and Adaptation The key observed and projected climate change impacts in Asia are summarized based on subsections 24.4.1 to 24.4.6 (Tables 24-1, 24-SM-4 and 24-SM-5). [INSERT TABLE 24-1 HERE Table 24-1: Key risks from climate change and the potential for risk reduction through mitigation and adaptation in Asia.] 24.4.1 Freshwater Resources 24.4.1.1. Sub-Regional Diversity Freshwater resources are very important in Asia because of the massive population and heavy economic dependence on agriculture, but water availability is highly uneven and requires assessment on the subregional scale because of Asia s huge range of climates (Pfister et al., 2009). Adequate water supply is one of the major challenges in many regions (Vörösmarty et al., 2010), particularly Central Asia. Growing demand for water is driven by soaring Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 populations, increasing per-capita domestic use, due to urbanization and thriving economic growth, and increasing use of irrigation. 24.4.1.2. Observed Impacts The impact of changes in climate, particularly precipitation, on water resources varies cross Asia (Table 24-SM-4). There is medium confidence that water scarcity in northern China has been exacerbated by decreasing precipitation, doubling population, and expanding water withdrawal from 1951 to 2000 (Xu et al., 2010). There is no evidence that suggests significant changes of groundwater in the Kherlen River Basin in Mongolia over the past half century (Brutsaert and Sugita, 2008). Apart from water availability, there is medium confidence that climate change also leads to degradation of water quality in most regions of Asia (Delpla et al., 2009; Park et al., 2010), although this is also heavily influenced by human activities (Winkel et al., 2011). Glaciers are important stores of water and any changes have the potential to influence downstream water supply in the long term (see Section 24.9.2). Glacier mass loss shows a heterogeneous pattern across Asia (Gardner et al., 2013). Glaciers in the polar section of the Ural Mountains, in the Kodar Mountains of Southeast Siberia, in the Suntar Khayata and Chersky Ranges of Northeast Siberia, in Georgia and Azerbaijan on the southern flank of the Greater Caucasus Range, on the Tibetan Plateau (see Box 3-1) and the surrounding areas, and on Puncak Jaya, Papua, Indonesia lost 9-80% of their total area in different periods within the 1895-2010 time interval (Ananicheva et al., 2005; Ananicheva et al., 2006; Anisimov et al., 2008; Prentice and Glidden, 2010; Allison, 2011; Shahgedanova et al., 2012; Yao et al., 2012a; Stokes et al., 2013) due to increased temperature (Casassa et al., 2009; Shrestha and Aryal, 2011). Changes in the Kamchatka glaciers are driven by both warming and volcanic activity, with the area of some glaciers decreasing, while others increased because they are covered by ash and clinker (Anisimov et al., 2008). 24.4.1.3. Projected Impacts Projected impacts of climate change on future water availability in Asia differ substantially among river basins and seasons (A1B scenario with 5 GCMs: Immerzeel et al., 2010; A1B with MRI-AGCMS: Nakaegawa et al., 2013) . There is high confidence that water demand in most Asian countries is increasing because of increases in population, irrigated agriculture (Lal, 2011) and industry. Tropical Asia. Future projections (A1B with MRI-AGCMs) suggest a decrease in river runoff in January in the Chao Phraya River basin in Thailand (Champathong et al., 2013). In a study of the Mahanadi River Basin in India, a water availability projection (A2, CGCM2) indicated increasing possibility of floods in September but increasing water scarcity in April (Asokan and Dutta, 2008). In the Ganges, an increase in river runoff could offset the large increases in water demand due to population growth in a +4C world (ensemble GCMs), due to a projected large increase in average rainfall, although high uncertainties remains at the seasonal scale (Fung et al., 2011). Northern and Temperate Asia. Projections (A2 and B2 with the GLASS model) suggest an increase in average water availability in Russia in the 2070s (Alcamo et al., 2007). In China, a projection (downscaling HadAM3H A2 and B2 scenarios with the PRECIS regional model) suggests that there will be insufficient water for agriculture in the 2020s and 2040s due to the increases in water demand for non-agricultural uses, although precipitation may increase in some areas (Xiong et al., 2010). In the late 21st century (MRI-AGCM, A1B), river discharge in northern Japan is projected to increase in February but decrease in May, due to increased winter precipitation and decreased spring snowmelt (Sato et al., 2013). Central and West Asia. Given the already very high level of water stress in many parts of Central Asia, projected temperature increases and precipitation decreases (SRES scenarios from IPCC AR4 23 models) in the western part of Kazakhstan, Uzbekistan, and Turkmenistan could exacerbate the problems of water shortage and distribution (Lioubimtseva and Henebry, 2009). Considering the dependence of Uzbekistan s economy on its irrigated Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 agriculture, which consumes more than 90% of the available water resources of the Amu Darya basin, climate change impacts on river flows would also strongly affect the economy (Schlüter et al., 2010). 24.4.1.4. Vulnerabilities to Key Drivers It is suggested that freshwater resource will be influenced by changes in rainfall variability, snowmelt or glacier retreat in the river catchment (Im et al., 2010; Ma et al., 2010; Sato et al., 2012; Yamanaka et al., 2012; Nakaegawa et al., 2013), and evapotranspiration, which are associated with climate change (Jian et al., 2009). Mismanagement of water resources has increased tension due to water scarcity in arid areas (Biswas and Seetharam, 2008; Lioubimtseva and Henebry, 2009; Siegfried et al., 2010; Aarnoudse et al., 2012). Unsustainable consumption of groundwater for irrigation and other uses is considered to be the main cause of groundwater depletion in the Indian states of Rajasthan, Punjab and Haryana (Rodell et al., 2009). 24.4.1.5. Adaptation Options Adaptation of freshwater resources to climate change can be identified as developing adaptive/integrated water resource management (Sadoff and Muller, 2009; Schlüter et al., 2010) of the trade-offs balancing water availability against increasing demand, in order to cope with uncertainty and change (Molle and Hoanh, 2009). Examples of the options include: developing water saving technologies in irrigation (Ngoundo et al., 2007); water infrastructure development in the Ganges river basin (Bharati et al., 2011); increasing water productivity in the Indus and Ganges river basins (Cai et al., 2010), Taiwan, China and the Philippines (Barker and Levine, 2012), and Uzbekistan (Tischbein et al., 2011); changing cropping systems and patterns in West Asia (Thomas, 2008); and water re-use in China (Yi et al., 2011). During the second half of the 20th century, Asia built many reservoirs and almost tripled its surface water withdrawals for irrigation. Reservoirs partly mitigate seasonal differences and increase water availability for irrigation (Biemans et al., 2011). Water management in river basins would benefit from integrated coordination among countries (Kranz et al., 2010). For example, water management in the Syr Darya river basin relates to Kyrgyzstan, Tajikistan, Uzbekistan, Turkmenistan, and Kazakhstan (Siegfried et al., 2010), while the Indus and Ganges-Brahmaputra-Meghna river basins concern Bangladesh, India, Nepal and Pakistan (Uprety and Salman, 2011). 24.4.2. Terrestrial and Inland Water Systems 24.4.2.1. Sub-Regional Diversity Boreal forests and grasslands dominate in North Asia, deserts and semi-deserts in Central and West Asia, and alpine ecosystems on the Tibetan Plateau. Human-dominated landscapes predominate in the other subregions, but the major natural ecosystems are temperate deciduous and subtropical evergreen forests in East Asia, with boreal forest in the northeast and grasslands and deserts in the west, while Southeast Asia was largely covered in tropical forests. South Asia also has tropical forests, with semi-desert in the northwest and alpine ecosystems in the north. Asia includes several of the world s largest river systems, as well as the world s deepest freshwater lake, Lake Baikal, the semi-saline Caspian Sea, and the saline Aral Sea. 24.4.2.2. Observed Impacts Biological changes consistent with climate trends have been reported in the north and at high altitudes, where rising temperatures have relaxed constraints on plant growth and the distributions of organisms. Few changes have been reported from tropical lowlands and none linked to climate change with high confidence, although data is insufficient to distinguish lack of observations from lack of impacts. Impacts on inland water systems have been difficult to disentangle from natural variability and other human impacts (Bates et al., 2008; Vörösmarty et al., 2010; Zheng, 2011; see Section 4.3.3.3). For example, the shrinking of the Aral Sea over the last 50 years has Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 resulted largely from excessive water extraction from rivers, but was probably exacerbated by decreasing precipitation and increasing temperature (Lioubimtseva and Henebry, 2009; Kostianoy and Kosarev, 2010). Phenology and Growth Rates. In humid temperate East Asia, plant observations and satellite measurements of greenness (Normalized Difference Vegetation Index, NDVI; see 4.3.2.2) show a trend to earlier leafing in spring since the 1980s, averaging 2 days a decade, although details vary between sites, species and periods (Table 24-SM-6) (detected with high confidence and attributed to warming with medium confidence.). Earlier spring flowering and delayed autumn senescence have also been recorded (Table 24-SM-6). Trends in semi-arid temperate regions were heterogeneous in space and time (Liu et al., 2013a; Yu et al., 2013a, 2013b). Earlier greening has been reported from boreal forests (Delbart et al., 2008) and from the Hindu-Kush-Himalayan region (Panday and Ghimire, 2012; Shrestha et al., 2012), but with spatial and temporal heterogeneity. Patterns were also heterogeneous in Central Asia (Kariyeva et al., 2012). On the Tibetan Plateau, spring growth advanced until the mid-1990s, but the trend subsequently differs between areas and NDVI datasets (Yu et al., 2010; Yu et al., 2012; Dong et al., 2013; Jin et al., 2013; Shen et al., 2013; Yu et al., 2013a; Zhang et al., 2013a; Zhang et al., 2013b). Satellite NDVI for Asia for 1988-2010 shows a general greening trend (i.e. increasing NDVI, a rough proxy for increasing plant growth), except where water is limiting (Dorigo et al., 2012). Changes at high latitudes (>60oN) show considerable spatial and temporal variability, despite a consistent warming trend, reflecting water availability and non-climatic factors (Bi et al., 2013; Jeong et al., 2013). Arctic tundra generally showed increased greening since 1982, while boreal forests were variable (Goetz et al., 2011; de Jong et al., 2012; Epstein et al., 2012; Xu et al., 2013). An overall greening trend for 2000-2011 north of the boreal forest correlated with increasing summer warmth and ice retreat (Dutrieux et al., 2012). In China, trends have varied in space and time, reflecting positive impacts of warming and negative impacts of increasing drought stress (Peng et al., 2011; Sun et al., 2012; Xu et al., 2012). The steppe region of northern Kazakhstan showed an overall browning (decreasing NDVI) trend for 1982-2008, linked to declining precipitation (de Jong et al., 2012). In Central Asia, where NDVI is most sensitive to precipitation (Gessner et al., 2013), there was a heterogeneous pattern for 1982-2009, with an initial greening trend stalled or reversed in some areas (Mohammat et al., 2013). Tree-ring data for 800-1989 for temperate East Asia suggests recent summer temperatures have exceeded those during past warm periods of similar length, although this difference was not statistically significant (Cook et al., 2012). Where temperature limits tree growth, growth rates have increased with warming in recent decades (Duan et al., 2010; Sano et al., 2010; Shishov and Vaganov, 2010; Borgaonkar et al., 2011; Xu et al., 2011; Li et al., 2012; Chen et al., 2012a, 2012b, 2012c, 2012d; Chen et al., 2013), while where drought limits growth, there have been increases (Li et al., 2006; Davi et al., 2009; Shao et al., 2010; Yang et al., 2010) or decreases (Li et al., 2007; Davi et al , 2009; Dulamsuren et al., 2010a, 2011; Kang et al., 2012; Wu et al., 2012; Kharuk et al., 2013; Liu et al., 2013b) reflecting decreasing or increasing water stress (high confidence in detection, medium confidence in attribution to climate change). In boreal forest, trends varied between species and locations, despite consistent warming (Lloyd and Bunn, 2007; Goetz et al., 2011). Distributions of Species and Biomes. Changes in species distributions consistent with a response to warming have been widely reported: upwards in elevation (Soja et al., 2007; Bickford et al., 2010; Kharuk et al., 2010a, 2010b, 2010e; Moiseev et al., 2010; Chen et al., 2011; Jump et al., 2012; Telwala et al., 2013; Grigor ev et al., 2013) or polewards (Tougou et al., 2009; Ogawa-Onishi and Berry, 2013) (high confidence in detection, medium confidence in attribution to climate change). Changes in the distributions of major vegetation types (biomes) have been reported from the north and high altitudes, where trees are invading treeless vegetation, and forest understories are being invaded from adjacent biomes (Soja et al., 2007; Kharuk et al., 2006; Bai et al., 2011; Singh et al., 2012; Wang and Liu, 2012; Ogawa-Onishi and Berry, 2013). In central Siberia, dark needle conifers (DNC) and birch have invaded larch-dominated forest over the last three decades (Kharuk et al., 2010c, d; Osawa et al., 2010; Lloyd et al., 2011). Meanwhile, warming has driven larch stand crown closure and larch invasion into tundra at a rate of 3 10 m/year in the northern forest-tundra ecotone (Kharuk et al., 2006). Shrub expansion in arctic tundra has also been observed (Blok et al., 2011; Myers-Smith et al., 2011; see 28.2.3.1.). Soil moisture and light are the main factors governing the forest-steppe ecotone (Soja et al., 2007; Zeng et al., 2008; Eichler et al., 2011; Kukavskaya et al., 2013) and Mongolian taiga forests have responded heterogeneously to recent climate changes, but declines in larch growth and regeneration are more widespread than increases (Dulamsuren et al., 2010a, 2010b). Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Permafrost. Permafrost degradation, including reduced area and increased active layer thickness, has been reported from parts of Siberia, Central Asia, and the Tibetan Plateau (Romanovsky et al., 2010; Wu and Zhang, 2010; Zhao et al., 2010; Yang et al., 2013) (high confidence). Most permafrost observatories in Asian Russia show substantial warming of permafrost during the last 20-30 years (Romanovsky et al., 2008, 2010). Permafrost formed during the Little Ice Age is thawing at many locations and Late Holocene permafrost has begun to thaw at some undisturbed locations in northwest Siberia. Permafrost thawing is most noticeable within the discontinuous permafrost zone, while continuous permafrost is starting to thaw in a few places, so the boundary between continuous and discontinuous permafrost is moving northwards (Romanovsky et al., 2008, 2010). Thawing permafrost may lead to increasing emissions of greenhouse gases from decomposition of accumulated organic matter (see Section 4.3.3.4 and 19.6.3.5). In Mongolia, mean annual permafrost temperature at 10-15 m depth increased over the past 10-40 years in the Hovsgol, Hangai and Hentei Mountain regions. Permafrost warming during the past 15 20 years was greater than during the previous 15-20 years (Sharkhuu et al., 2008; Zhao et al., 2010). In the Kazakh part of the Tien Shan Mountains, permafrost temperature and active layer thickness have increased since the early 1970s. Significant permafrost warming also occurred in the eastern Tien Shan Mountains, in the headwaters of the Urumqi River (Marchenko et al., 2007; Zhao et al., 2010). Monitoring across the Qinghai-Tibet Plateau over recent decades has also revealed permafrost degradation caused by warming and other impacts. Areas of permafrost are shrinking, the active layer depth is increasing, the lower altitudinal limit is rising, and the seasonal frost depth is thinning (Li et al., 2008; Wu and Zhang, 2010; Zhao et al., 2010). In the alpine headwater regions of the Yangtze and Yellow Rivers, rising temperatures and permafrost degradation have resulted in lower lake levels, drying swamps and shrinking grasslands (Cheng and Wu, 2007; Wang et al., 2011). 24.4.2.3. Projected Impacts Phenology and Growth Rates. Trends towards an earlier spring greening and longer growing season are expected to continue in humid temperate and boreal forest areas, although photoperiod or chilling requirements may reduce responses to warming in some species (Ge et al., 2013; Hadano et al., 2013; Richardson et al., 2013). Changes in precipitation will be important for semi-arid and arid ecosystems, as may the direct impacts of atmospheric CO2 concentrations, making responses harder to predict (Liancourt et al., 2012; Poulter et al., 2013). The general flowering at multi-year intervals in lowland rainforests in Southeast Asia is triggered by irregular droughts (Sakai et al., 2006), so changes in drought frequency or intensity could have large impacts. Distributions of Species and Biomes. Climate change is expected to modify the vegetation distribution across the region (Tao and Zhang, 2010; Wang, 2013), but responses will be slowed by limitations on seed dispersal, competition from established plants, rates of soil development, and habitat fragmentation (Corlett and Westcott, 2013) (high confidence). Rising CO2 concentrations are expected to favor increased woody vegetation in semi-arid areas (Higgins and Scheiter, 2012; Donohue et al., 2013; Poulter et al., 2013; Wang, 2013) (medium confidence). In North Asia, rising temperatures are expected to lead to large changes in the distribution of potential natural ecosystems (Ni, 2011; Tchebakova et al., 2011; Insarov et al., 2012; Pearson et al., 2013) (high confidence). It is likely that the boreal forest will expand northward and eastward, and that tundra will decrease, although differences in models, time periods, and other assumptions have resulted in widely varying projections for the magnitude of this change (Woodward and Lomas, 2004; Kaplan and New, 2006; Lucht et al., 2006; Golubyatnikov and Denisenko, 2007; Sitch et al., 2008; Korzukhin and Tcelniker, 2010; Tchebakova et al., 2010, 2011; Pearson et al., 2013). Boreal forest expansion and the continued invasion of the existing larch-dominated forest by dark-needle conifers could lead to larch reaching the Arctic shore, while the traditional area of larch dominance turns into mixed forest (Kharuk et al., 2006; Kharuk et al., 2010c). Both the replacement of summer-green larch with evergreen conifers and expansion of trees and shrubs into tundra decrease albedo, causing regional warming and potentially accelerating vegetation change (Kharuk et al., 2006; McGuire et al., 2007; Kharuk et al., 2010d; Pearson et al., 2013). The future direction and rate of change of steppe vegetation are unclear because of uncertain precipitation trends (Golubyatnikov and Denisenko, 2007; Tchebakova et al., 2010). The role of CO2-fertilization is also potentially important here (Poulter et al., 2013; see WG1 AR5 Box 6.3). Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In East Asia, subtropical evergreen forests are projected to expand north into the deciduous forest and tropical forests to expand along China s southern coast (Choi et al., 2011; Wang, 2013), but vegetation change may lag climate change by decades or centuries (Corlett and Westcott, 2013). On the Tibetan Plateau, projections suggest that alpine vegetation will be largely replaced by forest and shrubland, with tundra and steppe retreating to the north (Liang et al., 2012; Wang, 2013). Impacts in Central and West Asia will depend on changes in precipitation. In India, a dynamic vegetation model (A2 and B2 scenarios) projected changes in more than a third of the forest area by 2100, mostly from deciduous to evergreen forest in response to increasing rainfall, although fragmentation and other human pressures are expected to slow these changes (Chaturvedi et al., 2011). By 2100, large areas of tropical and subtropical lowland Asia are projected to experience combinations of temperature and rainfall outside the current global range, under a variety of model projections and emission scenarios (Williams et al., 2007; Beaumont et al., 2010; García-López and Allué, 2013), but the potential impacts of these novel conditions on biodiversity are largely unknown (Corlett, 2011). In Southeast Asia, projected climate (A2 and B1 scenarios) and vegetation changes are expected to produce widespread declines in bat species richness, northward range shifts for many species, and large reductions in the distributions of most species (Hughes et al., 2012). Projections for various bird species in Asia under a range of scenarios also suggest major impacts on distributions (Menon et al., 2009; Li et al., 2010; Ko et al., 2012). Projections for butterflies in Thailand (A2 and B2 scenarios) suggest that species richness within protected areas will decline c. 30% by 2070-2099 (Klorvuttimontara et al., 2011). Projections for dominant bamboos in the Qinling Mountains (A2 and B2 scenarios) suggest substantial range reductions by 2100, with potentially adverse consequences for the giant pandas which eat them (Tuanmu et al., 2012). Projections for snow leopard habitat in the Himalayas (B1, A1B and A2 scenarios) suggest contraction by up to 30% as forests replace open habitats (Forrest et al., 2012). Permafrost. In the Northern Hemisphere, a 20-90% decrease in permafrost area and a 50-300 cm increase in active layer thickness driven by surface warming is projected for 2100 by different models and scenarios (Schaefer et al., 2011). It is likely that permafrost degradation in North Asia will spread from the southern and low-altitude margins, advancing northwards and upwards, but rates of change vary greatly between model projections (Cheng and Wu, 2007; Riseborough et al., 2008; Romanovsky et al., 2008; Anisimov, 2009; Eliseev et al., 2009; Nadyozhina et al., 2010; Schaefer et al., 2011; Wei et al., 2011). Substantial retreat is also expected on the Qinghai-Tibet Plateau (Cheng and Wu, 2007). Near-surface permafrost is expected to remain only in Central and Eastern Siberia and parts of the QTP in the late 21st century. Inland Waters. Climate change impacts on inland waters will interact with dam construction, pollution, and land- use changes (Vörösmarty et al., 2010) (see also 24.9.1 and Section 3.3.2). Increases in water temperature will impact species and temperature-dependent processes (Hamilton, 2010; Dudgeon, 2011; Dudgeon, 2012). Coldwater fish will be threatened as rising water temperatures make much of their current habitat unsuitable (Yu et al., 2013c). Climate change is also expected to change flow regimes in running waters and consequently impact habitats and species that are sensitive to droughts and floods (see Box CC-RF). Habitats that depend on seasonal inundation, including floodplain grasslands and freshwater swamp forests, will be particularly vulnerable (Maxwell, 2009; Bezuijen, 2011; Arias et al., 2012). Reduced dry season flows are expected to combine with sea-level rise to increase saltwater intrusion in deltas (Hamilton, 2010; Dudgeon, 2012), although non-climatic impacts will continue to dominate in most estuaries (Syvitski et al., 2009). For most Asian lakes, it is difficult to disentangle the impacts of water pollution, hydro-engineering, and climate change (Battarbee et al., 2012). 24.4.2.4. Vulnerabilities to Key Drivers Permafrost melting in response to warming is expected to impact ecosystems across large areas (Cheng and Wu, 2007; Tchebakova et al., 2011) (high confidence). The biodiversity of isolated mountains may also be particularly vulnerable to warming, because many species already have small geographical ranges that will shrink further (La Sorte and Jetz, 2010; Liu et al., 2010; Chou et al., 2011; Noroozi et al., 2011; Peh et al., 2011; Jump et al., 2012; Tanaka et al., 2012a; Davydov et al., 2013). Many freshwater habitats are similarly isolated and their restricted- range species may be equally vulnerable (Dudgeon, 2012). In flatter topography, higher velocities of climate change Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (the speeds that species need to move to maintain constant climate conditions) increase the vulnerabilities of species that are unable to keep pace, as a result of limited dispersal ability, habitat fragmentation, or other non-climatic constraints (Corlett and Westcott, 2013). In the tropics, temperature extremes above the present range are a potential threat to organisms and ecosystems (Corlett, 2011; Jevanandam et al., 2013; Mumby et al., 2013). For much of interior Asia, increases in drought stress, as a result of declining rainfall and/or rising temperatures, are the key concern. Because aridity is projected to increase in the northern Mongolian forest belt during the 21st century (Sato et al., 2007), larch cover will likely be reduced (Dulamsuren et al., 2010a). In the boreal forest region, a longer, warmer growing season will increase vulnerability to fires, although other human influences may overshadow climate impacts in accessible areas (Flannigan et al., 2009; Liu et al., 2012; Li et al., 2013; see Section 4.3.3.1.1). If droughts intensify in lowland Southeast Asia, the synergies between warmth, drought, logging, fragmentation and fire (Daniau et al., 2012), and tree mortality (Kumagai and Porporato, 2012; Tan et al., 2013), possibly acerbated by feedbacks between deforestation, smoke aerosols and reduced rainfall (Aragao, 2012; Tosca et al., 2012), could greatly increase the vulnerability of fragmented forest landscapes (high confidence). 24.4.2.5. Adaptation Options Suggested strategies for maximizing the adaptive capacity of ecosystems include reducing non-climate impacts, maximizing landscape connectivity, and protecting refugia where climate change is expected to be less than the regional mean (Hannah, 2010; Game et al., 2011; Klorvuttimontara et al., 2011; Murthy et al., 2011; Ren et al., 2011; Shoo et al., 2011; Mandych et al., 2012). Additional options for inland waters include operating dams to maintain environmental flows for biodiversity, protecting catchments, and preserving river floodplains (Vörösmarty et al., 2010). Habitat restoration may facilitate species movements across climatic gradients (Klorvuttimontara et al., 2011; Hughes et al., 2012) and long-distance seed dispersal agents may need protection (McConkey et al., 2012). Assisted migration of genotypes and species is possible where movements are constrained by poor dispersal, but risks and benefits need to be considered carefully (Liu et al., 2010; Olden et al., 2010; Tchebakova et al., 2011; Dudgeon, 2012; Ishizuka and Goto, 2012; Corlett and Westcott, 2013). Ex situ conservation can provide back-up for populations and species most at risk from climate change (Chen et al., 2009). 24.4.3. Coastal Systems and Low-Lying Areas 24.4.3.1. Sub-Regional Diversity Asia s coastline includes the global range of shore types. Tropical and subtropical coasts support 45% of the world s mangrove forest (Giri et al., 2011) and low-lying areas in equatorial Southeast Asia support most of the world s peat swamp forests, as well as other forested swamp types. Intertidal salt marshes are widespread along temperate and arctic coasts, while a variety of non-forested wetlands occur inland. Asia supports 40% of the world s coral reef area, mostly in Southeast Asia, with the world s most diverse reef communities in the coral triangle (Spalding et al., 2001; Burke et al., 2011). Seagrass beds are widespread and support most of the world s seagrass species (Green and Short, 2003). Six of the seven species of sea turtle are found in the region and five nest on Asian beaches (Spotila, 2004). Kelp forests and other seaweed beds are important on temperate coasts (Bolton, 2010; Nagai et al., 2011). Arctic sea-ice supports a specialized community of mammals and other organisms (see Sections 28.2.3.3. and 28.2.3.4.). 24.4.3.2. Observed Impacts Most of Asia s non-Arctic coastal ecosystems are under such severe pressure from non-climate impacts that climate impacts are hard to detect (see Section 5.4.2). Most large deltas in Asia are sinking (as a result of groundwater withdrawal, floodplain engineering, and trapping of sediments by dams) much faster than global sea-level is rising (Syvitski et al., 2009). Widespread impacts can be attributed to climate change only for coral reefs, where the temporal and spatial patterns of bleaching correlate with higher than normal sea surface temperatures (see Section 5.4.2.4 and CC-CR) (very high confidence). Increased water temperatures may also explain declines in large Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 seaweed beds in temperate Japan (Nagai et al., 2011; see Section 5.4.2.3). Warming coastal waters have also been implicated in the northwards expansion of tropical and subtropical macroalgae and toxic phytoplankton (Nagai et al., 2011), fish (Tian et al., 2012), and tropical corals, including key reef-forming species (Yamano et al., 2011), over recent decades. The decline of large temperate seaweeds and expansion of tropical species in southwest Japan has been linked to rising sea surface temperatures (Tanaka et al., 2012b), and these changes have impacted fish communities (Terazono et al., 2012). In Arctic Asia, changes in permafrost and the effects of sea-level rise and sea-ice retreat on storm-wave energy have increased erosion (Are et al., 2008; Razumov, 2010; Handmer et al., 2012). Average erosion rates range from 0.27 m/year (Chukchi Sea) to 0.87 m/year (East Siberian Sea), with a number of segments in the Laptev and East Siberian Sea experiencing rates greater than 3 m/year (Lantuit et al., 2012). 24.4.3.3. Projected Impacts Marine biodiversity at temperate latitudes is expected to increase as temperature constraints on warm-water taxa are relaxed (see Section 6.4.1.1) (high confidence), but biodiversity in tropical regions may fall if, as evidence suggests, tropical species are already near their thermal maxima (Cheung et al., 2009, 2010; Nguyen et al., 2011) (medium confidence). Individual fish species are projected to shift their ranges northwards in response to rising sea surface temperatures (Tseng et al., 2011; Okunishi et al., 2012; Tian et al., 2012). The combined effects of changes in distribution, abundance and physiology may reduce the body size of marine fishes, particularly in the tropics and intermediate latitudes (Cheung et al., 2013). Continuation of current trends in sea-surface temperatures and ocean acidification would result in large declines in coral-dominated reefs by mid-century (Hoegh-Guldberg, 2011; Burke et al., 2011; see Section 5.4.2.4 and Box CC- CR) (high confidence). Warming would permit the expansion of coral habitats to the north but acidification is expected to limit this (Yara et al., 2012). Acidification is also expected to have negative impacts on other calcified marine organisms (algae, molluscs, larval echinoderms), while impacts on non-calcified species are unclear (Branch et al., 2013; Kroeker et al., 2013; See CC-OA). On rocky shores, warming and acidification are expected to lead to range shifts and changes in biodiversity (see Section 5.4.2.2). Future rates of sea-level rise are expected to exceed those of recent decades (see WGI AR5 Section 13.5.1), increasing coastal flooding, erosion, and saltwater intrusion into surface and groundwaters. In the absence of other impacts, coral reefs may grow fast enough to keep up with rising sea-levels (Brown et al., 2011; Villanoy et al., 2012; see Section 5.4.2.4), but beaches may erode and mangroves, salt marshes, and seagrass beds will decline, unless they receive sufficient fresh sediment to keep pace or they can move inland (Gilman et al., 2008; Bezuijen, 2011; Kintisch, 2013; see Section 5.3.2.3). Loucks et al. (2010) predict a 96% decline in tiger habitat in Bangladesh s Sunderbans mangroves with a 28 cm sea-level rise if sedimentation does not increase surface elevations. Rising winter temperatures are expected to result in poleward expansion of mangrove ecosystems (see Section 5.4.2.3). Coastal freshwater wetlands may be vulnerable to saltwater intrusion with rising sea-levels, but in most river deltas local subsidence for non-climatic reasons will be more important (Syvitski et al., 2009). Current trends in cyclone frequency and intensity are unclear (see 24.3.2 and Box CC-TC), but a combination of cyclone intensification and sea-level rise could increase coastal flooding (Knutson et al., 2010) and losses of coral reefs and mangrove forests would exacerbate wave damage (Gedan et al., 2011; Villanoy et al., 2012). In the Asian Arctic, rates of coastal erosion are expected to increase as a result of interactions between rising sea- levels and changes in permafrost and the length of the ice-free season (Pavlidis et al., 2007; Lantuit et al., 2012) (high agreement, medium evidence). The largest changes are expected for coasts composed of loose permafrost rocks and therefore subject to intensive thermal abrasion. If sea-level rises by 0.5 m over this century, modeling studies predict that the rate of recession will increase 1.5-2.6-fold for the coasts of the Laptev Sea, East Siberian Sea, and West Yamal in the Kara Sea, compared to the rate observed in the first years of the 21st century. Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 24.4.3.4. Vulnerabilities to Key Drivers Offshore marine systems are most vulnerable to rising water temperatures and ocean acidification, particularly for calcifying organisms such as corals. Sea-level rise will be the key issue for many coastal areas, particularly if combined with changes in cyclone frequency or intensity, or in Arctic Asia, with a lengthening open-water season. The expected continuing decline in the extent of sea-ice in the arctic may threaten the survival of some ice- associated organisms (see Section 28.2.2.1), with expanded human activities in previously inaccessible areas an additional concern (Post et al., 2013). 24.4.3.5. Adaptation Options The connectivity of marine habitats and dispersal abilities of marine organisms increase the capacity for autonomous (spontaneous) adaptation in coastal systems (Cheung et al., 2009). Creating marine protected areas where sea surface temperatures are projected to change least may increase their future resilience (Levy and Ban, 2013). For coral reefs, potential indicators of future resilience include later projected onset of annual bleaching conditions (van Hooidonk et al., 2013), past temperature variability, the abundance of heat-tolerant coral species, coral recruitment rates, connectivity, and macroalgae abundance (McClanahan et al., 2012). Similar strategies may help identify reefs that are more resilient to acidification (McLeod et al., 2013). Hard coastal defenses, such as sea walls, protect settlements at the cost of preventing adjustments by mangroves, salt marshes and seagrass beds to rising sea-levels. Landward buffer zones that provide an opportunity for future inland migration could mitigate this problem (Tobey et al., 2010). More generally, maintaining or restoring natural shorelines where possible is expected to provide coastal protection and other benefits (Tobey et al., 2010; Crooks et al., 2011). Projected increases in the navigability of the Arctic Ocean because of declining sea-ice suggest the need for a revision of environmental regulations in order to minimize the risk of marine pollution (Smith and Stephenson, 2013). 24.4.4. Food Production Systems and Food Security It is projected that climate change will affect food security by the middle of the 21st century, with the largest numbers of food-insecure people located in South Asia (see Chapter 7). 24.4.4.1. Sub-Regional Diversity AR4 Section 10.4.1.1 pointed out that there will be regional differences within Asia in the impacts of climate change on food production. Research since then has validated this divergence and new data are available especially for West and Central Asia (Tables 24-SM-4 and 24-SM-5). In AR4 Section 10.4.1, climate change was projected to lead mainly to reductions in crop yield. New research shows there will also be gains for specific regions and crops in given areas. Thus, the current assessment encompasses an enormous variability, depending on the regions and the crops grown. 24.4.4.2. Observed Impacts There are very limited data globally for observed impacts of climate change on food production systems (see Chapter 7) and this is true also for Asia. In Jordan, it was reported that the total production and average yield for wheat and barley were lowest in 1999 for the period 1996-2006 (Al-Bakri et al., 2010), which could be explained by the low rainfall during that year, which was 30% of the average (high confidence in detection, low confidence in attribution). In China, rice yield responses to recent climate change at experimental stations were assessed for the period 1981 2005 (Zhang et al., 2010). In some places, yields were positively correlated with temperature when they were also positively related with solar radiation. However, in other places, lower yield with higher temperature was accompanied by a positive correlation between yield and rainfall (high confidence in detection, high confidence in attribution). In Japan, where mean air temperature rose by about 1C over the 20th century, effects of recent warming include phenological changes in many crops, increases in fruit coloring disorders and incidences of chalky Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 rice kernels, reductions in yields of wheat, barley, vegetables, flowers, milk and eggs, and alterations in the type of disease and pest (high confidence in detection, high confidence in attribution) (Sugiura et al., 2012). 24.4.4.3. Projected Impacts Production. AR4 Section 10.4.1.1 mainly dealt with cereal crops (rice, wheat corn). Since then, impacts of climate change have been modeled for additional cereal crops and subregions. It is very likely that climate change effects on crop production in Asia will be variable, negative for specific regions and crops in given areas and positive for other regions and crops (high agreement, medium evidence). It is also likely that an elevated CO2 concentration in the atmosphere will be beneficial to most crops (high agreement, medium evidence). In semi-arid and arid regions of Western Asia, rainfed agriculture is sensitive to climate change both positively and negatively (Ratnakumar et al., 2011). In the mountainous Swat and Chitral districts of Pakistan (average altitudes 960 and 1500 m above sea level, respectively), there were mixed results as well (Hussain and Mudasser, 2007). Projected temperature increases of 1.5 and 3°C would lead to wheat yield declines (by 7% and 24% respectively) in Swat district but to increases (by 14% and 23%) in Chitral district. In India, climate change impacts on sorghum were analyzed using the InfoCrop-SORGHUM simulation model (Srivastava et al., 2010). A changing climate was projected to reduce monsoon sorghum grain yield by 2-14% by 2020, with worsening yields by 2050 and 2080. In the Indo-Gangetic Plains, a large reduction in wheat yields is projected (see below), unless appropriate cultivars and crop management practices are adopted (Ortiz et al., 2008). A systematic review and meta-analysis of data in 52 original publications projected mean changes in yield by the 2050s across South Asia of 16% for maize and 11% for sorghum (Knox et al., 2012). No mean change in yield was projected for rice. In China, modeling studies of the impacts of climate change on crop productivity have had mixed results. Rice is the most important staple food in Asia. Studies show that climate change will alter productivity in China but not always negatively. For example, an ensemble-based probabilistic projection shows rice yield in southeastern China would change on average by 7.5% to 17.5% ( 10.4% to 3.0%), 0.0% to 25.0% ( 26.7% to 2.1%), and 10.0% to 25.0% ( 39.2% to 6.4%) during the 2020s, 2050s, and 2080s, respectively, in response to climate change, with (without) consideration of CO2 fertilization effects, using all 10 combinations of two emission scenarios (A1FI and B1) and five GCMs (HadCM3, PCM, CGCM2, CSIRO2, and ECHAM4) relative to 1961 1990 levels (Tao and Zhang, 2013a). With rising temperatures, the process of rice development accelerates and reduces the duration for growth. Wassmann et al. (2009a, 2009b) concluded that, in terms of risks of increasing heat stress, there are parts of Asia where current temperatures are already approaching critical levels during the susceptible stages of the rice plant. These include: Pakistan/North India (October), South India (April, August), East India/Bangladesh (March-June), Myanmar/Thailand/Laos/Cambodia (March-June), Vietnam (April/August), Philippines (April/June), Indonesia (August) and China (July/August). There have also been simulation studies for other crops in China. In the Huang-Huai-Hai Plain, China s most productive wheat growing region, modeling indicated that winter wheat yields would increase on average by 0.2 Mg ha-1 in 2015-2045 and by 0.8 Mg ha-1 in 2070-2099, due to warmer nighttime temperatures and higher precipitation, under A2 and B2 scenarios using the HadCM3 model (Thomson et al., 2006). In the North China Plain, an ensemble-based probabilistic projection projected that maize yield will change by -9.7 to -9.1%, -19.0 to -15.7%, and -25.5% to -24.7%, during 2020s, 2050s, and 2080s as a percentage of 1961 1990 yields (Tao et al., 2009). In contrast, winter wheat yields could increase with high probability in future due to climate change (Tao and Zhang, 2013b). It should be noted that crop physiology simulation models such as those discussed above may overstate the impact of CO2 fertilization. Free atmosphere carbon exchange (FACE) experiments show that measurable CO2 fertilization effects are typically less than modeled results (see Section 7.3). Extreme weather events are also expected to negatively affect agricultural crop production (IPCC, 2012). For example, extreme temperatures could lower yields of rice (Mohammed and Tarpley, 2009; Tian et al., 2010). With higher precipitation, flooding could also lead to lower crop production (see SREX Chapter 4). Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Farming Systems and Crop Areas. Since the release of the AR4 (see WGII AR4 Section 10.4.1.2), more information is available on the impacts of climate change on farming systems and cropping areas in more countries in Asia and especially in Central Asia. Recent studies validate the likely northward shifts of crop production with current croplands under threat from the impacts of climate change (medium agreement, medium evidence). Cooler regions are likely to benefit as warmer temperatures increase arable areas (high agreement, medium evidence). Central Asia is expected to become warmer in the coming decades and increasingly arid, especially in the western parts of Turkmenistan, Uzbekistan, and Kazakhstan (Lioubimtseva and Henebry, 2009). Some parts of the region could be winners (cereal production in northern and eastern Kazakhstan could benefit from the longer growing season, warmer winters and a slight increase in winter precipitation), while others could be losers (particularly western Turkmenistan and Uzbekistan, where frequent droughts could negatively affect cotton production, increase already extremely high water demands for irrigation, and exacerbate the already existing water crisis and human- induced desertification). In India, the Indo-Gangetic Plains are under threat of a significant reduction in wheat yields (Ortiz et al., 2008). This area produces 90 million tons of wheat grain annually (about 14-15% of global wheat production). Climate projections based on a doubling of CO2 using a CCM3 model downscaled to a 30 arc-second resolution as part of the WorldClim data set showed that there will be a 51% decrease in the most favorable and high yielding area due to heat stress. About 200 million people (using the current population) in this area whose food intake relies on crop harvests would experience adverse impacts. Rice growing areas are also expected to shift with climate change throughout Asia. In Japan, increasing irrigation water temperature (1.6 2.0°C) could lead to a northward shift of the isochrones of safe transplanting dates for rice seedlings (Ohta and Kimura, 2007). As a result, rice cultivation period will be prolonged by approximately 25 30 days. This will allow greater flexibility in the cropping season than at present, resulting in a reduction in the frequency of cool-summer damage in the northern districts. Sea-level rise threatens coastal and deltaic rice production areas in Asia, such as those in Bangladesh and the Mekong River Delta (Wassmann et al., 2009b). For example, about 7% of Vietnam s agriculture land may be submerged due to sea-level rise (Dasgupta et al., 2009). In Myanmar, salt water intrusion due to sea-level rise could also decrease rice yield (Wassmann et al., 2009b). Fisheries and Aquaculture. Asia dominates both capture fisheries and aquaculture (FAO, 2010). More than half of the global marine fish catch in 2008 was in the West Pacific and Indian Ocean, and the lower Mekong River basin supports the largest freshwater capture fishery in the world (Dudgeon, 2011). Fish production is also a vital component of regional livelihoods, with 85.5% of the world s fishers (28 m) and fish farmers (10 m) in Asia in 2008. Many more people fish part-time. Fish catches in the Asian Arctic are relatively small, but important for local cultures and regional food security (Zeller et al., 2011). Inland fisheries will continue to be vulnerable to a wide range of on-going threats, including overfishing, habitat loss, water abstraction, drainage of wetlands, pollution, and dam construction, making the impacts of climate change hard to detect (see also 24.9.1). Most concerns have centered on rising water temperatures and the potential impacts of climate change on flow regimes, which in turn are expected to affect the reproduction of many fish species (Allison et al., 2009; Barange and Perry, 2009; Bezuijen, 2011; Dudgeon, 2011; see also Section 24.4.2.3). Sea-level rise is expected to impact both capture fisheries and aquaculture production in river deltas (De Silva and Soto, 2009). For marine capture fisheries, Cheung et al. (2009, 2010) used a dynamic bioclimate envelope model to project the distributions of 1066 species of exploited marine fish and invertebrates for 2005-2055, based on the SRES A1B scenario and a stable-2000 C02 scenario. This analysis suggests that climate change may lead to a massive redistribution of fisheries catch potential, with large increases in high-latitude regions, including Asian Russia, and large declines in the tropics, particularly Indonesia. Other studies have made generally similar predictions, with climate change impacts on marine productivity expected to be large and negative in the tropics, in part because of the vulnerability of coral reefs to both warming and ocean acidification (see also Section 24.4.3.3), and large and positive in arctic and subarctic regions, because of sea-ice retreat and poleward species shifts (Sumaila et al., 2011; Blanchard et al., 2012; Doney et al., 2012) (high confidence). Predictions of a reduction in the average maximum body weight of marine fishes by 14-24% by 2050 under a high-emission scenario are an additional threat to fisheries (Cheung et al., 2013). Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Future Food Supply and Demand. AR4 Section 10.4.1.4 was largely based on global models that included Asia. There are now a few quantitative studies in Asia and its individual countries. In general, these show that the risk of hunger, food insecurity and loss of livelihood due to climate change will likely increase in some regions (medium agreement, low evidence). Rice is a key staple crop in Asia and 90% or more of the world s rice production is from Asia. An Asia-wide study revealed that climate change scenarios (using 18 GCMs for A1B, 14 GCMs for A2, and 17 GCMs for B1) would reduce rice yield over a large portion of the continent (Masutomi et al., 2009). The most vulnerable regions were western Japan, eastern China, the southern part of the Indochina peninsula, and the northern part of South Asia. In Russia, climate change may also lead to food production shortfall , which was defined as an event in which the annual potential (i.e. climate-related) production of the most important crops in an administrative region in a specific year falls below 50% of its climate-normal (1961 1990) average (Alcamo et al., 2007). The study shows that the frequency of shortfalls in five or more of the main crop growing regions in the same year is around 2 years/decade under normal climate but could climb to 5 6 years/decade in the 2070s, depending on the scenario and climate model (using the GLASS and WaterGAP-2 models and ECHAM and HadCM3 under the A2 and B2 scenarios). The increasing shortfalls were attributed to severe droughts. The study estimated that the number of people living in regions that may experience one or more shortfalls each decade may grow to 82 139 million in the 2070s. Increasing frequency of extreme climate events will pose an increasing threat to the security of Russia s food system. In contrast, climate change may provide a windfall for wheat farmers in parts of Pakistan. Warming temperatures would make it possible to grow at least two crops (wheat and maize) a year in mountainous areas (Hussain and Mudasser, 2007). In the northern mountainous areas, wheat yield was projected to increase by 50% under SRES A2 and by 40% under the B2 scenario, whereas in the sub-mountainous, semi-arid and arid areas, it is likely to decrease, by the 2080s (Iqbal et al., 2009). 24.4.4.4. Vulnerabilities to Key Drivers Food production and food security are most vulnerable to rising air temperatures (Wassmann et al., 2009a, 2009b). Warmer temperatures could depress yields of major crops such as rice. However, warmer temperatures could also make some areas more favorable for food production (Lioubimtseva and Henebry, 2009). Increasing CO2 concentration in the atmosphere could lead to higher crop yields (Tao and Zhang, 2013a). Sea-level rise will be a key issue for many coastal areas as rich agricultural lands may be submerged and taken out of production (Wassmann et al., 2009b). 24.4.4.5. Adaptation Options Since AR4, there have been additional studies of recommended and potential adaptation strategies and practices in Asia (Table 24-SM-7) and there is new information for West and Central Asia. There are also many more crop- specific and country-specific adaptation options available. Farmers have been adapting to climate risks for generations. Indigenous and local adaptation strategies have been documented for Southeast Asia (Peras et al., 2008; Lasco et al., 2010; Lasco et al., 2011) and could be used as a basis for future climate change adaptation. Crop breeding for high temperature condition is a promising option for climate change adaptation in Asia. For example, in the North China Plain simulation studies show that using high-temperature sensitive varieties, maize yield in the 2050s could increase on average by 1.0 6.0%, 9.9 15.2%, and 4.1 5.6%, by adopting adaptation options of early planting, xing variety growing duration, and late planting, respectively (Tao and Zhang, 2010). In contrast, no adaptation will result in yield declines of 13.2 19.1%. Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 24.4.5. Human Settlements, Industry, and Infrastructure 24.4.5.1. Sub-Regional Diversity Around one in every five urban dwellers in Asia lives in large urban agglomerations and almost 50% of these live in small cities (UN, 2012). North and Central Asia are the most urbanized areas, with over 63% of the population living in urban areas, with the exception of Kyrgyzstan and Tajikistan (UN-Habitat, 2010; UN ESCAP, 2011). South and Southwest Asia are the least urbanized subregions, with only a third of their populations living in urban areas. However, these regions have the highest urban population growth rates within Asia at an average of 2.4% per year during 2005-2010 (UN ESCAP, 2011). By the middle of this century, Asia s urban population will increase by 1.4 billion and will account for over 50% of the global population (UN, 2012). 24.4.5.2. Observed Impacts Asia experienced the highest number of weather- and climate-related disasters in the world during the period 2000- 2008 and suffered huge economic losses, accounting for the second highest proportion (27.5%) of the total global economic loss (IPCC, 2012). Flood mortality risk is heavily concentrated in Asia. Severe floods in Mumbai in 2005 have been attributed to both climatic factors and non-climatic factors. Strengthened capacities to address the mortality risk associated with major weather-related hazards, such as floods, have resulted in a downward trend in mortality risk relative to population size, as in East Asia, where it is now a third of its 1980 level (UNISDR, 2011). 24.4.5.3. Projected Impacts A large proportion of Asia s population lives in low elevation coastal zones that are particularly at risk from climate change hazards, including sea-level rise, storm surges and typhoons (see Sections 5.3.2.1 and 8.2.2.5, Box CC-TC). Depending on the region, half to two-thirds of Asia s cities with 1 million or more inhabitants are exposed to one or multiple hazards, with floods and cyclones the most important (UN, 2012). Floodplains and Coastal Areas. Three of the world s five most populated cities (Tokyo, Delhi and Shanghai) are located in areas with high risk of floods (UN, 2012). Flood risk and associated human and material losses are heavily concentrated in India, Bangladesh, and China. At the same time, the East Asia region in particular is experiencing increasing water shortages, negatively affecting its socioeconomic, agricultural, and environmental conditions, which is attributed to lack of rains and high evapotranspiration, as well as over-exploitation of water resources (IPCC, 2012). Large parts of South, East and Southeast Asia are exposed to a high degree of cumulative climate-related risk (UN-Habitat, 2011). Asia has more than 90% of the global population exposed to tropical cyclones (IPCC, 2012); see Box CC-TC). Damage due to storm surge is sensitive to change in the magnitude of tropical cyclones. By the 2070s, the top Asian cities in terms of population exposure (including all environmental and socioeconomic factors) to coastal flooding are expected to be Kolkata, Mumbai, Dhaka, Guangzhou, Ho Chi Minh City, Shanghai, Bangkok, Rangoon, and Hai Phong (Hanson et al., 2011). The top Asian cities in terms of assets exposed are expected to be Guangdong, Kolkata, Shanghai, Mumbai, Tianjin, Tokyo, Hong Kong, and Bangkok. Asia includes 15 of the global top 20 cities for projected population exposure and 13 of the top 20 for asset exposure. Other Issues in Human Settlements. Asia has a large and rapidly expanding proportion of the global urban exposure and vulnerability related to climate change hazards (see SREX Section 4.4.3). In line with the rapid urban growth and sprawl in many parts of Asia, the periurban interface between urban and rural areas deserves particular attention when considering climate change vulnerability (see also Section 18.4.1). Garschagen et al. (2011) find, for example, that periurban agriculturalists in the Vietnamese Mekong Delta are facing a multiple burden since they are often exposed to overlapping risks resulting from (a) socio-economic transformations, such as land title insecurity and price pressures; (b) local biophysical degradation, as periurban areas serve as sinks for urban wastes; and (c) climate change impacts, as they do not benefit from the inner-urban disaster risk management measures. Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Nevertheless, the periurban interface is still underemphasized in studies on impacts, vulnerability and adaptation in Asia. Groundwater sources, which are affordable means of high-quality water supply in cities of developing countries, are threatened due to over-withdrawals. Aquifer levels have fallen by 20-50 m in cities such as Bangkok, Manila and Tianjin and between 10-20 m in many other cities (UNESCO, 2012). The drop in groundwater levels often results in land subsidence, which can enhance hazard exposure due to coastal inundation and sea-level rise, especially in settlements near the coast, and deterioration of groundwater quality. Cities susceptible to human-induced subsidence (mainly, developing country cities in deltaic regions with rapidly growing populations) could see significant increases in exposure (Nicholls et al., 2008). Settlements on unstable slopes or landslide-prone areas face increased prospects of rainfall-induced landslides (IPCC, 2012). Industry and Infrastructure. The impacts of climate change on industry include both direct impacts on industrial production and indirect impacts on industrial enterprises due to the implementation of mitigation activities (Li, 2008). The impact of climate change on infrastructure deterioration cannot be ignored, but can be addressed by changes to design procedures, including increases in cover thickness, improved quality of concrete, and coatings and barriers (Stewart et al., 2012). Climate change and extreme events may have a greater impact on large and medium- sized construction projects (Kim et al., 2007). Estimates suggest that by upgrading the drainage system in Mumbai, losses associated with a 1-in-100 year flood event today could be reduced by as much as 70%, and through extending insurance to 100% penetration, the indirect effects of flooding could be almost halved, speeding recovery significantly (Ranger et al., 2011). On the east coast of India, clusters of districts with poor infrastructure and demographic development are also the regions of maximum vulnerability. Hence, extreme events are expected to be more catastrophic in nature for the people living in these districts. Moreover, the lower the district is in terms of the infrastructure index and its growth, the more vulnerable it is to the potential damage from extreme events and hence people living in these regions are prone to be highly vulnerable (Patnaik and Narayanan, 2009). In 2008, the embankments on the Kosi River (a tributary of the Ganges) failed, displacing over sixty thousand people in Nepal and three and a half million in India. Transport and power systems were disrupted across large areas. However, the embankment failure was not caused by an extreme event but represented a failure of interlinked physical and institutional infrastructure systems in an area characterized by complex social, political, and environmental relationships (Moench, 2010). 24.4.5.4. Vulnerabilities to Key Drivers Disruption of basic services such as water supply, sanitation, energy provision, and transportation systems have implications for local economies and strip populations of their assets and livelihoods , in some cases leading to mass migration (UN-Habitat, 2010). Such impacts are not expected to be evenly spread among regions and cities, across sectors of the economy, or among socioeconomic groups. They tend to reinforce existing inequalities and disrupt the social fabric of cities and exacerbate poverty. 24.4.5.5. Adaptation Options An ADB and UN report estimates that about two-thirds of the $8 trillion needed for infrastructure investment in Asia and the Pacific between 2010 and 2020 will be in the form of new infrastructure, which creates tremendous opportunities to design, finance and manage more sustainable infrastructure (ADB et al., 2012).) Adaptation measures that offer a no regrets solution are proposed for developing countries, where basic urban infrastructure is often absent (e.g. appropriate drainage infrastructure), leaving room for actions that both increase immediate well- being and reduce vulnerability to future climate change (Hallegatte and Corfee-Morlot, 2011). The role of urban planning and urban planners in adaptation to climate change impacts has been emphasized (Fuchs et al., 2011; IPCC, 2012; Tyler and Moench, 2012). The focus on solely adapting through physical infrastructure in urban areas requires complementary adapting planning, management, governance and institutional arrangements to be able to deal with Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 the uncertainty and the unprecedented challenges implied by climate change (Revi, 2008; Birkmann et al., 2010; Garschagen and Kraas, 2011). 24.4.6. Human Health, Security, Livelihoods, and Poverty 24.4.6.1. Sub-Regional Diversity Although rapidly urbanizing, Asia is still predominantly an agrarian society, with 57.28% of its total population living in rural areas, of which 81.02% are dependent on agriculture for their livelihoods (FAOSTAT, 2011). Rural poverty is higher than urban poverty, reflecting the heavy dependence on natural resources that are directly influenced by changes in weather and climate (Haggblade et al., 2010; IFAD, 2010). Rural poverty is expected to remain more prevalent than urban poverty for decades to come (Ravallion et al., 2007). However, climate change will also affect urbanizing Asia, where the urban poor will be impacted indirectly, as evident from the food price rises in the Middle East and other areas in 2007-2008. Certain categories of urban dwellers, such as urban wage labor households, are particularly vulnerable (Hertel et al., 2010). Agriculture has been identified as a key driver of economic growth in Asia (World Bank, 2007). Although economic growth was impressive in recent decades, there are still gaps in development compared to the rest of the world (World Bank, 2011). Southeast Asia is the third poorest performing region after Sub-Saharan Africa and Southern Asia in terms of the Human Development Indicators (UN, 2009). Impacts on human security in Asia will primarily manifest through impacts on water resources, agriculture, coastal areas, resource-dependent livelihoods, and urban settlements and infrastructure, with implications for human health and well-being. Regional disparities on account of socioeconomic context and geographical characteristics largely define the differential vulnerabilities and impacts within countries in Asia (Thomas, 2008; Sivakumar and Stefanski, 2011). 24.4.6.2. Observed Impacts Floods and Health. Epidemics have been reported after floods and storms (Bagchi, 2007) as a result of decreased drinking water quality (Harris et al., 2008; Hashizume et al., 2008; Solberg, 2010; Kazama et al., 2012), mosquito proliferation (Pawar et al., 2008), and exposure to rodent-borne pathogens (Kawaguchi et al., 2008; Zhou et al., 2011) and the intermediate snail hosts of Schistosoma (Wu et al., 2008). Contaminated urban flood waters have caused exposure to pathogens and toxic compounds, for example in India and Pakistan (Sohan et al., 2008; Warraich et al., 2011). Mental disorders and posttraumatic stress syndrome have also been observed in disaster prone areas (Udomratn, 2008) and, in India, have been linked to age and gender (Telles et al., 2009). See also Chapter 11.4.2. for flood-attributable deaths. Heat and Health. The effects of heat on mortality and morbidity have been studied in many countries, with a focus on the elderly and people with cardiovascular and respiratory disorders (Kan et al., 2007; Guo et al., 2009; Huang et al., 2010). Associations between high temperatures and mortality have been shown for populations in India and Thailand (McMichael et al., 2008) and in several cities in East Asia (Kim et al., 2006; Chung et al., 2009). Several studies have analyzed the health effects of air pollution in combination with increased temperatures (Lee et al., 2007; Qian et al., 2010; Wong et al., 2010; Yi et al., 2010). Intense heat waves have been shown to affect outdoor workers in South and East Asia (Nag et al., 2007; Hyatt et al., 2010). Drought and Health. Dust storms in Southwest, Central and East Asia result in increased hospital admissions and worsen asthmatic conditions, as well as causing skin and eye irritations (Griffin, 2007; Hashizume et al., 2010; Kan et al., 2012). Droughts may also lead to wildfires and smoke exposure, with increased morbidity and mortality, as observed in Southeast Asia (Johnston et al., 2012). Drought can also disrupt food security, increasing malnutrition (Kumar et al., 2005) and thus susceptibility to infectious diseases. Water-borne Diseases. Many pathogens and parasites multiply faster at higher temperatures. Temperature increases have been correlated with increased incidence of diarrheal diseases in East Asia (Huang et al., 2008; Zhang et al., Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2008; Onozuka et al., 2010). Other studies from South and East Asia have shown an association between diarrheal outbreaks and a combination of higher temperatures and heavy rainfall (Hashizume et al., 2007; Majra and Gur, 2009; Chou et al., 2010). Increasing coastal water temperatures correlated with outbreaks of systemic Vibrio vulnificus infection in Israel (Paz et al., 2007) and South Korea (Kim and Jang, 2010). Cholera outbreaks in coastal populations in South Asia have been associated with increased water temperatures and algal blooms (Huq et al., 2005). The ENSO cycle and Indian Ocean Dipole have been associated with cholera epidemics in Bangladesh (Pascual, 2000; Rodó et al., 2002; Hashizume et al., 2011). Vector-borne Diseases. Increasing temperatures affect vector-borne pathogens during the extrinsic incubation period and shorten vector life-cycles, facilitating larger vector populations and enhanced disease transmission, whilst the vector s ability to acquire and maintain a pathogen tails off (Paaijmans et al., 2012). Dengue outbreaks in South and Southeast Asia are correlated with temperature and rainfall with varying time lags (Su, 2008; Hii et al., 2009; Hsieh and Chen, 2009; Shang et al., 2010; Sriprom et al., 2010; Hashizume et al., 2012). Outbreaks of vaccine- preventable Japanese encephalitis have been linked to rainfall in studies from the Himalayan region (Partridge et al., 2007; Bhattachan et al., 2009), and to rainfall and temperature in South and East Asia (Bi et al., 2007; Murty et al., 2010). Malaria prevalence is often influenced by non-climate variability factors, but studies from India and Nepal have found correlations with rainfall (Devi and Jauhari, 2006; Dev and Dash, 2007; Dahal, 2008; Laneri et al., 2010). Temperature was linked to distribution and seasonality of malaria mosquitoes in Saudi Arabia (Kheir et al., 2010). The re-emergence of malaria in central China has been attributed to rainfall and increases in temperature close to water bodies (Zhou et al., 2010). In China, temperature, precipitation, and the virus-carrying index among rodents have been found to correlate with the prevalence of hemorrhagic fever with renal syndrome (Guan et al., 2009). Livelihoods and Poverty. An estimated 51% of total income in rural Asia comes from non-farm sources (Haggblade et al., 2009, 2010), mostly local non-farm business and employment. The contribution of remittances to rural income has grown steadily (Estudillo and Otsuka, 2010). Significant improvements have been made in poverty eradication over the past decade (World Bank, 2008), with rapid reductions in poverty in East Asia, followed by South Asia (IFAD, 2010). A significant part of the reduction has come from population shifts, rapid growth in agriculture, and urban contributions (Janvry and Sadoulet, 2010). Climate change negatively impacts livelihoods (see Table 24-SM- 4) and these impacts are directly related to natural resources affected by changes in weather and climate. Factors that have made agriculture less sustainable in the past include input non-responsive yields, soil erosion, natural calamities, and water and land quality related problems (Dev, 2011). These have predisposed rural livelihoods to climate change vulnerability. Livelihoods are impacted by droughts (Harshita, 2013; Selvaraju et al., 2006), floods (Nuorteva et al., 2010; Dun, 2011; Nguyen, 2007; Keskinen et al., 2010) and typhoons (Huigen and Jens, 2006; Gaillard et al., 2007; Uy et al., 2011). Drought disproportionately impacts small farmers, agricultural laborers, and small businessmen (Selvaraju et al., 2006), who also have least access to rural safety net mechanisms, including financial services (IFAD, 2010), despite recent developments in microfinance services in parts of Asia. Past floods have exposed conditions such as lack of access to alternative livelihoods, difficulty in maintaining existing livelihoods, and household debts leading to migration in the Mekong region (Dun, 2011). Similar impacts of repeated floods leading to perpetual vulnerability were found in the Tonle Sap Lake area of Cambodia (Nuorteva et al., 2010; Keskinen et al., 2010). Typhoon impacts are mainly through damage to the livelihood assets of coastal populations in the Philippines and the level of ownership of livelihood assets has been a major determinant of vulnerability (Uy et al., 2011). 24.4.6.3. Projected Impacts Health Effects. An emerging public health concern in Asia is increasing mortality and morbidity due to heat waves. An ageing population will increase the number of people at risk, especially those with cardiovascular and respiratory disorders. Urban heat island effects have increased (Tan et al., 2010), although local adaptation of the built environment and urban planning will determine the impacts on public health. Heat stress disorders among workers and consequent productivity losses have also been reported (Lin et al., 2009; Langkulsen et al., 2010). The relationship between temperature and mortality is often U-shaped (Guo et al., 2009), with increased mortality also during cold events, particularly in rural environments, even if temperatures do not fall below 0°C (Hashizume et al., Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2009). However, some studies in developing areas suggest that factors other than climate can be important, so warming may not decrease cold-related deaths much in these regions (Honda and Ono, 2009). Climate change will affect the local transmission of many climate-sensitive diseases. Increases in heavy rain and temperature are projected to increase the risk of diarrheal diseases in, for example, China (Zhang et al., 2008). However, the impact of climate change on malaria risk will differ between areas, as projected for West and South Asia (Husain and Chaudhary, 2008; Garg et al., 2009; Majra and Gur, 2009), while a study suggested that the impact of socioeconomic development will be larger than that of climate change (Béguin et al., 2011). Climate change is also expected to affect the spatiotemporal distribution of dengue fever in the region, although the level of evidence differs across geographical locations (Banu et al., 2011). Some studies have developed climate change-disease prevalence models, for example one for schistosomiasis in China shows an increased northern distribution of the disease with climate change (Zhou et al., 2008; Kan et al., 2012). Impacts of climate change on fish production (Qiu et al., 2010) are being studied, along with impacts on chemical pathways in the marine environment and consequent impacts on food safety (Tirado et al., 2010), including seafood safety (Marques et al., 2010). Livelihood and Poverty. Floods, droughts and changes in seasonal rainfall patterns are expected to negatively impact crop yields, food security and livelihoods in vulnerable areas (Dawe et al., 2008; Kelkar et al., 2008; Douglas, 2009). Rural poverty in parts of Asia could be exacerbated (Skoufias et al., 2011) due to impacts on the rice crop and increases in food prices and the cost of living (Hertel et al., 2010; Rosegrant, 2011). The poverty impacts of climate change will be heterogeneous among countries and social groups (see Table 24-SM-5). In a low crop productivity scenario, producers in food exporting countries, such as Indonesia, the Philippines and Thailand, would benefit from global food price rises and reduce poverty, while countries such as Bangladesh would experience a net increase in poverty of 15% by 2030 (Hertel et al., 2010). These impacts will also differ within food exporting countries, with disproportionate negative impacts on farm laborers and the urban poor. Skoufias et al. (2011) project significant negative impacts of a rainfall shortfall on the welfare of rice farmers in Indonesia, compared to a delay in rainfall onset. These impacts may lead to global mass migration and related conflicts (Laczko and Aghazarm, 2009; Barnett and Webber, 2010; Warner, 2010; World Bank, 2010). In North Asia, climate-driven changes in tundra and forest-tundra biomes may influence indigenous peoples who depend on nomadic tundra pastoralism, fishing and hunting (Kumpula et al., 2011). 24.4.6.4. Vulnerabilities to Key Drivers Key vulnerabilities vary widely within the region. Climate change can exacerbate current socio-economic and political disparities and add to the vulnerability of Southeast Asia and Central Asia to security threats that may be transnational in nature (Jasparro and Taylor, 2008; Lioubimtseva and Henebry, 2009). Apart from detrimental impacts of extreme events, vulnerability of livelihoods in agrarian communities also arises from geographic settings, demographic trends, socio-economic factors, access to resources and markets, unsustainable water consumption, farming practices and lack of adaptive capacity (Mulligan et al., 2011; Acosta-Michlik and Espaldon, 2008; Allison et al., 2009; Knox et al., 2011; Lioubimtseva and Henebry, 2009; Byg and Salick, 2009; Salick and Ross, 2009; Salick et al., 2009; Xu et al., 2009; UN, 2009). Urban wage laborers were found to be more vulnerable to cost of living related poverty impacts of climate change than those who directly depend on agriculture for their livelihoods (Hertel et al., 2010). In Indonesia, drought-associated fires increase vulnerability of agriculture, forestry and human settlements, particularly in peatland areas (Murdiyarso and Lebel, 2007). Human health is also a major area of focus for Asia (Munslow and O'Dempsey, 2010), where the magnitude and type of health effects from climate change depend on differences in socioeconomic and demographic factors, health systems, the natural and built environment, land use changes, and migration, in relation to local resilience and adaptive capacity. The role of institutions is also critical, particularly in influencing vulnerabilities arising from gender (Ahmed and Fajber, 2009), caste and ethnic differences (Jones and Boyd, 2011), and securing climate-sensitive livelihoods in rural areas (Agrawal and Perrin, 2008). Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 24.4.6.5. Adaptation Options Disaster preparedness on a local community level could include a combination of indigenous coping strategies, early-warning systems, and adaptive measures (Paul and Routray, 2010). Heat warning systems have been successful in preventing deaths among risk groups in Shanghai (Tan et al., 2007). New work practices to avoid heat stress among outdoor workers, in Japan and the UAE have also been successful (Morioka et al., 2006; Joubert et al., 2011). Early warning models have been developed for haze exposure from wildfires, in for example Thailand (Kim Oanh and Leelasakultum, 2011), and are being tested in infectious disease prevention and vector control programs, as for malaria in Bhutan (Wangdi et al., 2010) and Iran (Haghdoost et al., 2008), or are being developed, as for dengue fever region-wide (Wilder-Smith et al., 2012). Some adaptation practices provide unexpected livelihood benefits, as with the introduction of traditional flood mitigation measures in China which could positively impact local livelihoods, leading to reductions in both the physical and economic vulnerabilities of communities (Xu et al., 2009). A greater role of local communities in decision making is also proposed (Alauddin and Quiggin, 2008) and in prioritization and adoption of adaptation options (Prabhakar et al., 2010; Prabhakar and Srinivasan, 2011). Defining adequate community property rights, reducing income disparity, exploring market-based and off-farm livelihood options, moving from production-based approaches to productivity and efficiency decision-making based approaches, and promoting integrated decision- making approaches, have also been suggested (Merrey et al., 2005; Brouwer et al., 2007; Paul et al., 2009; Niino, 2011; Stucki and Smith, 2011). Climate resilient livelihoods can be fostered through the creation of bundles of capitals (natural, physical, human, financial and social capital) and poverty eradication (Table 24-SM-8). Greater emphasis on agricultural growth has been suggested as an effective means of reducing rural poverty (Janvry and Sadoulet, 2010; Rosegrant, 2011). Bundled approaches are known to facilitate better adaptation than individual adaptation options (Acosta-Michlik and Espaldon, 2008; Fleischer et al., 2011). Community-based approaches have been suggested to identify adaptation options that address poverty and livelihoods, as these techniques help capture information at the grassroots (Huq and Reid, 2007; van Aalst et al., 2008), and help integration of disaster risk reduction, development, and climate change adaptation (Heltberg et al., 2010), connect local communities and outsiders (van Aalst et al., 2008), address the location-specific nature of adaptation (Iwasaki et al., 2009; Rosegrant, 2011), help facilitate community learning processes (Bass and Ramasamy, 2008), and help design location-specific solutions (Ensor and Berger, 2009). Some groups can become more vulnerable to change after being locked into specialized livelihood patterns, as with fish farmers in India (Coulthard, 2008). Livelihood diversification, including livelihood assets and skills, has been suggested as an important adaptation option for buffering climate change impacts on certain kinds of livelihoods (Selvaraju et al., 2006; Nguyen, 2007; Agrawal and Perrin, 2008; IFAD, 2011; Keskinen et al., 2010; Uy et al., 2011). The diversification should occur across assets, including productive assets, consumption strategies and employment opportunities (Agrawal and Perrin, 2008). Ecosystem-based adaptation has been suggested to secure livelihoods in the face of climate change (Jones et al., 2012), integrating the use of biodiversity and ecosystem services into an overall strategy to help people adapt (IUCN, 2009). Among financial means, low-risk liquidity options such as microfinance programs and risk transfer products can help lift the rural poor from poverty and accumulate assets (Barrett et al., 2007; Jarvis et al., 2011). 24.4.7 Valuation of Impacts and Adaptation Economic valuation in Asia generally covers impacts and vulnerabilities of disperse sectors such as food production, water resources and human health (Aydinalp and Cresser, 2008; Kelkar et al., 2008; Lioubimtseva and Henebry, 2009; Su et al., 2009; Srivastava et al., 2010). Multi-sector evaluation that unpacks the relationships between and across sectors, particularly in a context of resource scarcity and competition, is very limited. Information is scarce especially for North, Central and West Asia. Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Generally, annual losses from drought are expected to increase based on various projection under diverse scenarios, but such losses are expected to be reduced if adaptation measures are implemented (ADB, 2009; Sutton et al., 2013). It is also stressed that there are great uncertainties associated with the economic aspects of climate change. In China, the total loss due to drought projected in 2030 is expected to range from $1.1-1.7 billion for regions in northeast China and about $0.9 billion for regions in north China (CWF et al., 2009), with adaptation measures having the potential to avert half of the losses. In India, the estimated countrywide agricultural loss in 2030 of over $7 billion that will severely affect the income of 10% of the population could be reduced by 80% if cost-effective climate resilience measures are implemented (CWF et al., 2009). In Indonesia, the Philippines, Thailand and Vietnam, under the A2 scenario, the PAGE2002 integrated assessment model projects a mean loss of 2.2% of gross domestic product (GDP) by 2100 on an annual basis, if only the market impact (mainly related to agriculture and coastal zones) is considered (ADB, 2009). This is well above the world s projected mean GDP loss of 0.6% each year by 2100 due to market impact alone. In addition, the mean cost for the four countries could reach 5.7% of the GDP if non-market impacts related to health and ecosystems are included and 6.7% of the GDP if catastrophic risks are also taken into account. The cost of adaptation for agriculture and coastal zones is expected to be about $5 billion/year by 2020 on average. Adaptation that is complemented with global mitigation measures is expected to be more effective in reducing the impacts of climate change (IPCC, 2007; ADB, 2009; UNFCCC, 2009; MNRE, 2010; Begum et al., 2011). 24.5. Adaptation and Managing Risks 24.5.1. Conservation of Natural Resources Natural resources are already under severe pressure from land-use change and other impacts in much of Asia. Deforestation in Southeast Asia has received most attention (Sodhi et al., 2010; Miettinen et al., 2011a), but ecosystem degradation, with the resulting loss of natural goods and services, is also a major problem in other ecosystems. Land-use change is also a major source of regional greenhouse gas emissions, particularly in Southeast Asia (see WGI AR5 Section 6.3.2.2 and Table 6.3). Projected climate change is expected to intensify these pressures in many areas (see Sections 24.4.2.3 and 24.4.3.3), most clearly for coral reefs, where increases in sea surface temperature and ocean acidification are a threat to all reefs in the region and the millions of people who depend on them (see Section 5.4.2.4 and Boxes CC-CR and CC-OA). Adaptation has so far focused on minimizing non-climate pressures on natural resources and restoring connectivity to allow movements of genes and species between fragmented populations (see Section 24.4.2.5). Authors have also suggested a need to identify and protect areas that will be subject to the least damaging climate change ( climate refugia ) and to identify additions to the protected area network that will allow for expected range shifts, for example by extending protection to higher altitudes or latitudes. Beyond the intrinsic value of wild species and ecosystems, ecosystem-based approaches to adaptation aim to use the resilience of natural systems to buffer human systems against climate change, with potential social, economic and cultural co-benefits for local communities (see Box CC-EA). 24.5.2. Flood Risks and Coastal Inundation Many coasts in Asia are exposed to threats from floods and coastal inundation (see also 24.4.5.3). Responding to a large number of climate change impact studies for each Asian country over the past decade (e.g. Karim and Mimura, 2008; Pal and Al-Tabbaa, 2009), various downscaled tools to support, formulate and implement climate change adaptation policy for local governments are under development. One of the major tools is vulnerability assessment and policy option identification with Geographical Information Systems (GIS). These tools are expected to be of assistance in assessing city-specific adaptation options by examining estimated impacts and identified vulnerability for some coastal cities and areas in Asian countries (e.g. Brouwer et al., 2007; Taylor, 2011; Storch and Downs, 2011). These tools and systems sometimes take the form of integration of top down approaches and bottom-up (community-based) approaches (see Section 14.5). Whereas top-down approaches give scientific knowledge to local actors, community-based approaches are built on existing knowledge and expertise to strengthen coping and adaptive capacity by involving local actors (van Aalst et al., 2008). Community-based approaches may have a Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 limitation in that they place greater responsibility on the shoulders of local people without necessarily increasing their capacity proportionately (Allen, 2006). As the nature of adaptive capacity varies depending on the formulation of social capital and institutional context in the local community, it is essential for the approaches to be based on an understanding of local community structures (Adger, 2003). 24.5.3. Economic Growth and Equitable Development Climate change challenges fundamental elements in social and economic policy goals such as prosperity, growth, equity and sustainable development (Mearns and Norton, 2010). Economic, social, and environmental equity is an enduring challenge in many parts of Asia. Generally, the level of wealth (typically GDP) has been used as a measure of human vulnerability of a country but this approach has serious limitations (Mattoo and Subramanian, 2012; Dellink et al., 2009). In many cases, social capital, an indicator of equity in income distribution within countries, is a more important factor in vulnerability and resilience than GDP per capita (Lioubimtseva and Henebry, 2009; Islam et al., 2006). Furthermore, political and institutional instabilities can undermine the influence of economic development (Lioubimtseva and Henebry, 2009). Poor and vulnerable countries are at greater risk of inequity and loss of livelihoods from the impacts of climate extremes as their options for coping with such events are limited. Many factors contribute to this limitation, including poverty, illiteracy, weak institutions and infrastructures, poor access to resources, information and technology, poor health care, and low investment and management capabilities. The overexploitation of land resources including forests, increases in population, desertification and land degradation pose additional threats (UNDP, 2006). This is particularly true for developing countries in Asia with a high level of natural resource dependency. Provision of adequate resources based on the burden sharing and the equity principle will serve to strengthen appropriate adaptation policies and measures in such countries (Su et al., 2009). 24.5.4. Mainstreaming and Institutional Barriers Mainstreaming climate change adaptation into sustainable development policies offers a potential opportunity for good practice to build resilience and reduce vulnerability, depending on effective, equitable and legitimate actions to overcome barriers and limits to adaptation (ADB, 2005; Lim et al., 2005; Lioubimtseva and Henebry, 2009). The level of adaptation mainstreaming is most advanced in the context of official development assistance, where donor agencies and international financial institutions have made significant steps towards taking climate change adaptation into account in their loan and grant making processes (Gigli and Agrawala, 2007; Klein et al., 2007). While some practical experiences of adaptation in Asia at the regional, national and local level are emerging, there can be barriers that impede or limit adaptation. These include challenges related to competing national priorities, awareness and capacity, financial resources for adaptation implementation, institutional barriers, biophysical limits to ecosystem adaptation, and social and cultural factors (Lasco et al., 2009; Moser and Ekstrom, 2010; Lasco et al., 2012). Issues with resource availability might not only result from climate change, but also from weak governance mechanisms and the breakdown of policy and regulatory structures, especially with common-pool resources (Moser and Ekstrom, 2010). Furthermore, the impact of climate change depends on the inherent vulnerability of the socio- ecological systems in a region as much as on the magnitude of the change (Evans, 2010). Recent studies linking climate-related resource scarcities and conflict call for enhanced regional cooperation (Gautam, 2012). 24.5.5. Role of Higher Education in Adaptation and Risk Management To enhance the development of young professionals in the field of climate change adaptation, the topic could be included in higher education, especially in formal education programs. Shaw et al. (2011) mentioned that higher education in adaptation and disaster risk reduction in the Asia-Pacific region can be done through environment disaster linkage, focus on hydro-meteorological disasters, and emphasizing synergy issues between adaptation and risk reduction. Similar issues are also highlighted by other authors (Chhokar, 2010; Niu et al., 2010; Nomura and Abe, 2010; Ryan et al., 2010). Higher education should be done through lectures and course work, field studies, internships, and establishing education-research link by exposing the students to field realities. In this regard, Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 guiding principles could include: an inclusive curriculum, focus on basic theory, field orientation, multidisciplinary courses and practical skill enhancement. Bilateral or multilateral practical research programs on adaptation and risk management by the graduate students and young faculty members would expose them to the real field problems. 24.6. Adaptation and Mitigation Interactions Integrated mitigation and adaptation responses focus on either land-use changes or technology development and use. Changes in land use, such as agroforestry, may provide both mitigation and adaptation benefits (Verchot et al., 2007), or otherwise, depending on how they are implemented. Agroforestry practices provide carbon storage and may decrease soil erosion, increase resilience against floods, landslides and drought, increase soil organic matter, reduce the financial impact of crop failure, as well as have biodiversity benefits over other forms of agriculture, as shown, for example, in Indonesia (Clough et al., 2011). Integrated approaches are often needed when developing mitigation-adaptation synergies, as seen in waste-to-compost projects in Bangladesh (Ayers and Huq, 2009). Other adaptation measures that increase biomass and/or soil carbon content, such as ecosystem protection and reforestation, will also contribute to climate mitigation by carbon sequestration. However, exotic monocultures may fix more carbon than native mixtures while supporting less biodiversity and contributing less to ecological services, calling for compromises that favor biodiversity-rich carbon storage (Diaz et al., 2009). The potential for both adaptation and mitigation through forest restoration is greatest in the tropics (Sasaki et al., 2011). At higher latitudes (>45oN), reforestation can have a net warming influence by reducing surface albedo (Anderson-Teixeira et al., 2012). Expansion of biofuel crops on abandoned and marginal agricultural lands could potentially make a large contribution to mitigation of carbon emissions from fossil fuels, but could also have large negative consequences for both carbon and biodiversity if it results directly or indirectly in the conversion of carbon-rich ecosystems to cropland (Fargione et al., 2010; Qin et al., 2011). Mechanisms, such as REDD+, that put an economic price on land- use emissions, could reduce the risks of such negative consequences (Thomson et al., 2010), but the incentive structures need to be worked out very carefully (Busch et al., 2012). Forests and their management are also often emphasized for providing resilient livelihoods and reducing poverty (Chhatre and Agrawal, 2009; Noordwijk, 2010; Persha et al., 2010; Larson, 2011). Securing rights to resources is essential for greater livelihood benefits for poor indigenous and traditional people (Macchi et al., 2008) and the need for REDD+ schemes to respect and promote community forest tenure rights has been emphasized (Angelsen, 2009). It has been suggested that indigenous people can provide a bridge between biodiversity protection and climate change adaptation (Salick and Ross, 2009): a point that appears to be missing in the current discourse on ecosystem- based adaptation. There are arguments against REDD+ supporting poverty reduction due to its inability to promote productive use of forests, which may keep communities in perpetual poverty (Campbell, 2009), but there is a contrasting view that REDD+ can work in forests managed for timber production (Putz et al., 2012; Guariguata et al., 2008), especially through reduced impact logging (Guariguata et al., 2008) and other approaches such as assuring the legality of forest products, certifying responsible management, and devolving control over forests to empowered local communities (Putz et al., 2012). On rivers and coasts, the use of hard defenses (e.g. sea-walls, channelization, bunds, dams) to protect agriculture and human settlements from flooding may have negative consequences for both natural ecosystems and carbon sequestration by preventing natural adjustments to changing conditions (see 24.4.3.5). Conversely, setting aside landward buffer zones along coasts and rivers would be positive for both. The very high carbon sequestration potential of the organic-rich soils in mangroves (Donato et al., 2011) and peat swamp forests (Page et al., 2011) provides opportunities for combining adaptation with mitigation through restoration of degraded areas. Mitigation measures can also result in public health benefits (Bogner et al., 2008; Haines et al., 2009). For example, sustainable cities with fewer fossil-fuel driven vehicles (mitigation) and more trees and greenery (carbon storage and adaptation to the urban heat island effect) would have a number of co-benefits, including public health a promising strategy for triple win interventions (Romero-Lankao et al., 2011). Other examples include efforts to decarbonize electricity production in India and China that are projected to decrease mortality due to reduced PM5 and PM2.5 particulates (Markandya et al., 2009); policies to increase public transportation, promote walking and cycling, and reduce private cars that will increase air quality and decrease the health burden, particularly in urban environments Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 as projected in India (Woodcock et al., 2009) ; and abandoning the use of biomass fuel or coal for indoor cooking and heating to improve indoor air quality and respiratory and cardiac health among, in particular, women and children in India and China (Wilkinson et al., 2009). Conversely, actions to reduce current environmental-public health issues may often have beneficial mitigation effects, like traffic emissions reduction programs in China (Wu et al., 2011) and India (Reynolds and Kandikar, 2008). 24.7. Intra-regional and Inter-regional Issues 24.7.1. Trans-boundary Pollution Many Asian countries and regions face long-distance and trans-boundary air pollution problems. In eastern China, Japan and the Korean Peninsula, these include dust storms that originate in the arid and semi-arid regions upwind, with impacts on climate, human health and ecosystems (Huang et al., 2013). The susceptibility of the land surface to wind erosion is strongly influenced by vegetation cover, which is in turn sensitive to climate change and other human impacts. In the humid tropics of Southeast Asia, in contrast, the major trans-boundary pollution issue involves smoke aerosols from burning of biomass and peatlands, mostly during clearance for agriculture (Miettinen et al., 2011b; Gautam et al., 2013). Apart from the large impact on human health, these aerosols may be having a significant effect on rainfall in equatorial regions, leading to the possibility of climate-feedbacks, with fires reducing rainfall and promoting further fires (Tosca et al., 2012). Pollutants of industrial origin are also a huge problem in many parts of the region, with well-documented impacts on human health (see Section 24.4.6) and the climate (see WGI AR5 Chapters 7 and 8). 24.7.2. Trade and Economy The ASEAN Free Trade Agreement (AFTA) and the Indonesia Japan Economic Partnership Agreement (IJEPA) have positively impacted the Indonesian economy and reduced water pollution, but increased CO2 emissions by 0.46% compared to the business-as-usual situation, mainly due to large emission increases in the transportation sector (Gumilang et al., 2011). Full liberalization of tariffs and GDP growth concentrated in China and India has led to transport emissions growing much faster than the value of trade, due to a shift towards distant trading partners (Cristea et al., 2013). China's high economic growth and flourishing domestic and international trade has resulted in increased consumption and pollution of water resources (Guan and Hubacek, 2007). Japanese imports from the ASEAN region are negatively correlated with per capita carbon emissions (Atici, 2012) due to strict regulations in Japan that prevent import from polluting sectors. Export-led growth is central to the economic progress and well- being of Southeast Asian countries. Generally, as exports rise, carbon emissions tend to rise. International trading systems that help address the challenge of climate change need further investigation. 24.7.3. Migration and Population Displacement Floods and droughts are predominant causes for internal displacement (Internal Displacement Monitoring Center, 2011). In 2010 alone, 38.3 million people were internally displaced; 85% because of hydrological hazards and 77% in Asia. Floods are increasingly playing a role in migration in the Mekong Delta (Warner, 2010). Often some migrants return to the vulnerable areas (Piguet, 2008) giving rise to ownership, rights of use, and other issues (Kolmannskog, 2008). Increasing migration has led to increasing migration-induced remittances contributing to Asian economies, but has had negligible effect on the poverty rate (Vargas-Silva et al., 2009). In Bangladesh, migrant workers live and work under poor conditions, such as crowded shelters, inadequate sanitation, conflict and competition with the local population, and exploitation (Penning-Rowsell et al., 2011). Forced migration can result from adaptation options such as construction of dams, but the negative outcomes could be allayed by putting proper safeguards in place (Penning-Rowsell et al., 2011). Managed retreat of coastal communities is a suggested option to address projected sea-level rise (Alexander et al., 2012). A favorable approach to deal with migration is within a development framework and through adaptation strategies (Penning-Rowsell et al., 2011; ADB, 2012). Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 24.8. Research and Data Gaps Studies of observed climate changes and their impacts are still inadequate for many areas, particularly in North, Central and West Asia (Table 24-2). Improved projections for precipitation, and thus water supply, are most urgently needed. Another priority is developing water management strategies for adaptation to changes in demand and supply. More research is also needed on the health effects of changes in water quality and quantity. Understanding of climate change impacts on ecosystems and biodiversity in Asia is currently limited by the poor quality and low accessibility of biodiversity information (UNEP, 2012). National biodiversity inventories are incomplete and few sites have the baseline information needed to identify changes. For the tropics, major research gaps include the temperature dependence of carbon fixation by tropical trees, the thermal tolerances and acclimation capacities of both plants and animals, and the direct impacts of rising CO2 (Corlett, 2011; Zuidema et al., 2013). Rising CO2 is also expected to be important in cool-arid ecosystems, where lack of experimental studies currently limits our ability to make predictions (Poulter et al., 2013). Boreal forest dynamics will be influenced by complex interactions between rising temperatures and CO2, permafrost thawing, forest fires, and insect outbreaks (Osawa et al., 2010; Zhang et al., 2011), and understanding this complexity will require enhanced monitoring of biodiversity and species ranges, improved modeling, and greater knowledge of species biology (Meleshko and Semenov, 2008). Rice is the most studied crop but there are still significant uncertainties in model accuracy, CO2-fertilization effects, and regional differences (Masutomi et al., 2009; Zhang et al., 2010; Shuang-He et al., 2011). For other crops, there is even greater uncertainty. Studies are also needed of the health effects of interactions between heat and air pollution in urban and rural environments. More generally, research is needed on impacts, vulnerability and adaptation in urban settlements, especially cities with populations under 500,000, which share half the region s urban population. Greater understanding is required of the linkages between local livelihoods, ecosystem functions, and land resources for creating a positive impact on livelihoods in areas with greater dependence on natural resources (Paul et al., 2009). Increasing regional collaboration in scientific research and policy making has been suggested for reducing climate change impacts on water, biodiversity and livelihoods in the Himalayan region (Xu et al., 2009) and could be considered elsewhere. The literature suggests that work must begin now on building understanding of the impacts of climate change and moving forward with the most cost-effective adaptation measures (ADB, 2007; Cai et al., 2008; Mathy and Guivarch, 2010; Stage, 2010). For devising mitigation policies, the key information needed is again the most cost-effective measures (Nguyen, 2007; Cai et al., 2008; Mathy and Guivarch, 2010). [INSERT TABLE 24-2 HERE Table 24-2: The amount of information supporting conclusions regarding observed and projected impacts in Asia. 24.9. Case Studies 24.9.1. Transboundary Adaptation Planning and Management Lower Mekong River Basin The Lower Mekong River Basin (LMB) covers an area of approximately 606,000 sq. km across the countries of Thailand, Laos, Cambodia and Vietnam. More than 60 million people are heavily reliant on natural resources, in particular agriculture and fisheries, for their well-being (MRC, 2009; Dugan et al., 2010; Figure 24-SM-2). Thailand and Vietnam produced 51% of the world s rice exports in 2008, mostly in the LMB (Mainuddin et al., 2011). Observations of climate change over the past 30-50 years in the LMB include: an increase in temperature, an increase in rainfall in the wet season and decreases in the dry season, intensified flood and drought events, and sea- level rise (ICEM, 2010; IRG, 2010). Agricultural output has been noticeably impacted by intensified floods and droughts which caused almost 90% of rice production losses in Cambodia during 1996-2001 (Brooks and Adger, 2003; MRC, 2009).Vietnam and Cambodia are two of the countries most vulnerable to climate impacts on fisheries (Allison et al., 2009; Halls, 2009). Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Existing studies about future climate impacts in the Mekong Basin broadly share a set of common themes (MRC, 2009; Murphy and Sampson, 2013): increased temperature and annual precipitation; increased depth and duration of flood in the Mekong Delta and Cambodia floodplain; prolonged agricultural drought in the south and the east of the basin; and sea-level rise and salinity intrusion in the Mekong delta. Hydropower dams along the Mekong River and its tributaries will also have severe impacts on fish productivity and biodiversity, by blocking critical fish migration routes, altering the habitat of non-migratory fish species, and reducing nutrient flows downstream (Costanza et al., 2011; Baran and Guerin, 2012; Ziv et al., 2012). Climate impacts, though less severe than the impact of dams, will exacerbate these changes (Wyatt and Baird, 2007; Grumbine et al., 2012; Orr et al., 2012; Räsänen et al., 2012; Ziv et al., 2012). National climate change adaptation plans have been formulated in all four LMB countries, but transboundary adaptation planning across the LMB does not exist to date. Effective future transboundary adaptation planning and management will benefit from: a shared climate projection across the LMB for transboundary adaptation planning; improved coordination among adaptation stakeholders and sharing of best practices across countries; mainstreaming climate change adaptation into national and sub-national development plans with proper translation from national adaptation strategies into local action plans; integration of transboundary policy recommendations into national climate change plans and policies; integration of adaptation strategies on a landscape scale between ministries and different levels of government within a country (MRC, 2009; Lian and Bhullar, 2011; Lebel et al., 2012; Kranz et al., 2010) A study of the state-of-adaptation practice in the LMB showed that only 11% (45 of 417) of climate-change related projects in the LMB were on-the-ground adaptation efforts driven by climate risks (Ding, 2012; Neo, 2012; Schaffer and Ding, 2012). Common features of successful projects include: robust initial gap assessment, engagement of local stakeholders, and a participatory process throughout (Brown, 2012; Khim, 2012; Mondal, 2012; Panyakul, 2012; Roth and Grunbuhel, 2012). A multi-stakeholder Regional Adaptation Action Network has been proposed with the intent of scaling up and improving mainstreaming of adaptation through tangible actions following the theory and successful examples of the Global Action Networks (GANs) (Waddell, 2005; Waddell and Khagram, 2007; WCD, 2000; GAVI, 2011; Schaffer and Ding, 2012) . 24.9.2. Glaciers of Central Asia In the late 20th century, central Asian glaciers occupied 31,628 km2 (Dolgushin and Osipova). All recent basin-scale studies document multidecadal area loss (Figure 24-3); where multiple surveys are available, most show accelerating loss. The rate of glacier area change varies (Table 24-SM-9). Rates between 0.05%/yr and 0.76%/yr have been reported in the Altai (Surazakov et al., 2007; Shahgedanova et al., 2010; Yao et al., 2012b) and Tien Shan (Lettenmaier et al., 2009; Sorg et al., 2012), and between 0.13%/yr and 0.30%/yr in the Pamir (Konovalov and Desinov, 2007; Aizen, 2011a, 2011b, 2011c; Yao et al., 2012b). These ranges reflect varying sub-regional distributions of glacier size (smaller glaciers shrink faster) and debris cover (which retards shrinkage), but also varying proportions of ice at high altitudes, where as yet warming has produced little increase in melt (Narama et al., 2010). Most studies also document mean-annual (e.g. Glazyrin and Tadzhibaeva, 2011, for 1961-1990) and summertime (e.g. Shahgedanova et al., 2010) warming, with slight cooling in the central and eastern Pamir (Aizen, 2011b). Precipitation increases have been observed more often than decreases (e.g. Braun et al., 2009; Glazyrin and Tadzhibaeva, 2011). [INSERT FIGURE 24-3 HERE Figure 24-3: Losses of glacier area in the Altai-Sayan, Pamir and Tien Shan. Remote sensing data analysis from 1960s (Corona) through 2008 (Landsat, ASTER and Alos Prism).] Aizen et al. (2007) calculated 21st-century losses of 43% of the volume of Tien Shan glaciers for an 8oC temperature increase accompanied by a 24% precipitation increase, but probable complete disappearance of glaciers if precipitation decreased by 16%; a more moderate 2oC increase led to little loss, but only if accompanied by a 24% precipitation increase. Drawing on CMIP5 simulations, (Radiæ et al., 2013) simulated losses by 2100 of between 25% and 90% of 2006 ice volume (including Tibet but excluding the Altai and Sayan; range of all single-model Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 simulations); the 14-GCM model mean losses are 55% for RCP4.5 and 75% for RCP8.5. Similarly, Marzeion et al. (2012) found 21st-century volume losses of 50% for RCP2.6, about 57% for both RCP4.5 and RCP6.0, and 67% for RCP8.5. The glaciers have therefore been a diminishing store of water, and the diminution is projected to continue. Paradoxically, this implies more meltwater, possibly explaining limited observations of increased runoff (Sorg et al., 2012), but also an eventual decrease of meltwater yield (see Section 3.4.4). More immediately, it entails a hazard due to the formation of moraine-dammed glacial lakes (Bolch et al., 2011). Frequently Asked Questions FAQ 24.1: What will the projected impact of future climate change be on freshwater resources in Asia? [to be placed in Section 24.4.1] Asia is a huge and diverse region, so both climate change and the impact on freshwater resources will vary greatly depending on location. But throughout the region, adequate water resources are particularly important because of the massive population and heavy dependence of the agricultural sector on precipitation, river runoff and groundwater. Overall, there is low confidence in the projections of specifically how climate change will impact future precipitation on a subregional scale, and thus in projections of how climate change might impact the availability of water resources. However, water scarcity is expected to be a big challenge in many Asian regions because of increasing water demand from population growth and consumption per capita with higher standards of living. Shrinkage of glaciers in central Asia is expected to increase due to climate warming, which will influence downstream river runoff in these regions. Better water management strategies could help ease water scarcity. Examples include developing water saving technologies in irrigation, building reservoirs, increasing water productivity, changing cropping systems and water reuse. FAQ 24.2: How will climate change affect food production and food security in Asia? [to be placed in Section 24.4.4] Climate change impacts on temperature and precipitation will affect food production and food security in various ways in specific areas throughout this diverse region. Climate change will have a generally negative impact on crop production Asia, but with diverse possible outcomes [medium confidence]. For example most simulation models show that higher temperatures will lead to lower rice yields as a result of a shorter growing period. But some studies indicate that increased atmospheric CO2 that leads to those higher temperatures could enhance photosynthesis and increase rice yields. This uncertainty on the overall effects of climate change and CO2 fertilization is generally true for other important food crops such as wheat, sorghum, barley, and maize among others. Yields of some crops will increase in some areas (e.g. cereal production in north and east Kazakhstan) and decrease in others (e.g. wheat in the Indo-Gangetic Plain of South Asia). In Russia, climate change may lead to a food production shortfall, defined as an event in which the annual potential production of the most important crops falls 50% or more below it's normal average. Sea-level rise is projected to decrease total arable areas and thus food supply in many parts of Asia. A diverse mix of potential adaptation strategies, such as crop breeding, changing crop varieties, adjusting planting time, water management, diversification of crops and a host of indigenous practices will all be applicable within local contexts. FAQ 24.3: Who is most at risk from climate change in Asia? [to be placed in Section 24.4.6] People living in low-lying coastal zones and flood plains are probably most at risk from climate change impacts in Asia. Half of Asia's urban population lives in these areas. Compounding the risk for coastal communities, Asia has more than 90% of the global population exposed to tropical cyclones. The impact of such storms, even if their frequency or severity remains the same, is magnified for low lying and coastal zone communities because of rising sea level [medium confidence]. Vulnerability of many island populations is also increasing due to climate change impacts. Settlements on unstable slopes or landslide prone-areas, common in some parts of Asia, face increased likelihood of rainfall-induced landslides. Asia is predominantly agrarian, with 58% of its population living in rural areas, of which 81% are dependent on agriculture for their livelihoods. Rural poverty in parts of Asia could be exacerbated due to negative impacts from climate change on rice production, and a general increase in food prices and the cost of living [high confidence]. Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climate change will have widespread and diverse health impacts. More frequent and intense heatwaves will increase mortality and morbidity in vulnerable groups in urban areas [high confidence]. The transmission of infectious disease, such as cholera epidemics in coastal Bangladesh, and schistosomiasis in inland lakes in China, and diarrheal outbreaks in rural children will be affected due to warmer air and water temperatures and altered rain patterns and water flows [medium confidence]. Outbreaks of vaccine-preventable Japanese encephalitis in the Himalayan region and malaria in India and Nepal have been linked to rainfall. Changes in the geographical distribution of vector-borne diseases, as vector species that carry and transmit diseases migrate to more hospitable environments, will occur [medium confidence]. These effects will be most noted close to the edges of the current habitats of these species. Cross-Chapter Box Box CC-TC. Building Long-Term Resilience from Tropical Cyclone Disasters [Yoshiki Saito (Japan), Kathleen McInnes (Australia)] Tropical cyclones (also referred to as hurricanes and typhoons in some regions or strength) cause powerful winds, torrential rains, high waves and storm surge, all of which can have major impacts on society and ecosystems. Bangladesh and India account for 86% of mortality from tropical cyclones (Murray et al., 2012), which is mainly due to the rarest and most severe storm categories (i.e. Categories 3, 4, and 5 on the Saffir-Simpson scale). About 90 tropical cyclones occur globally each year (Seneviratne et al., 2012) although interannual variability is large. Changes in observing techniques particularly after the introduction of satellites in the late 1970s, confounds the assessment of trends in tropical cyclone frequencies and intensities. Therefore, IPCC (2012) Special Report on Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX) concluded that there is low confidence that any observed long-term (i.e. 40 years or more) increases in tropical cyclone activity are robust, after accounting for past changes in observing capability (Seneviratne et al., 2012; Chapter 2). There is also low confidence in the detection and attribution of century scale trends in tropical cyclones. Future changes to tropical cyclones arising from climate change are likely to vary by region. This is because there is medium confidence that for certain regions, shorter-term forcing by natural and anthropogenic aerosols has had a measurable effect on tropical cyclones. Tropical cyclone frequency is likely to decrease or remain unchanged over the 21st century, while intensity (i.e. maximum wind speed and rainfall rates) is likely to increase (AR5 WG1 Ch 14.6). Regionally specific projections have lower confidence (see AR5 WG1 Box 14.2). Longer-term impacts from tropical cyclones include salinisation of coastal soils and water supplies and subsequent food and water security issues from the associated storm surge and waves (Terry and Chui, 2012). However, preparation for extreme tropical cyclone events through improved governance and development to reduce their impacts provides an avenue for building resilience to longer-term changes associated with climate change. Densely populated Asian deltas are particularly vulnerable to tropical cyclones due to their large population density in expanding urban areas (Nicholls et al., 2007). Extreme cyclones in Asia since 1970 caused over 0.5 million fatalities (Murray et al., 2012) e.g., cyclones Bhola in 1970, Gorky in 1991, Thelma in 1998, Gujarat in 1998, Orissa in 1999, Sidr in 2007, and Nargis in 2008. Tropical cyclone Nargis hit Myanmar on 2 May 2008 and caused over 138,000 fatalities. Several-meter high storm surges widely flooded densely populated coastal areas of the Irrawaddy Delta and surrounding areas (Revenga et al., 2003; Brakenridge et al., 2013). The flooded areas were captured by a NASA MODIS image on 5 May 2008 (see Figure TC-1). [INSERT FIGURE TC-1 HERE Figure TC-1: The intersection of inland and storm surge flooding. Red shows May 5, 2008 MODIS mapping of the tropical cyclone Nargis storm surge along the Irrawaddy Delta and to the east, Myanmar. The blue areas to the north were flooded by the river in prior years. Source: Brakenridge et al., 2013.] Murray et al. (2012) compared the response to cyclone Sidr in Bangladesh in 2007 and Nargis in Myanmar in 2008 and demonstrated how disaster risk reduction methods could be successfully applied to climate change adaptation. Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sidr, despite being of similar strength to Nargis, caused far fewer fatalities (3,400 compared to over 138000) and this was attributed to advancement in preparedness and response in Bangladesh through experience in previous cyclones such as Bhola and Gorky. The responses included the construction of multistoried cyclone shelters, improvement of forecasting and warning capacity, establishing a coastal volunteer network, and coastal reforestation of mangroves. The strategies of disaster risk management for tropical cyclones in coastal areas, that create protective measures, anticipate and plan for extreme events, increase the resilience of potentially exposed communities. The integration of activities relating to education, training, and awareness-raising into relevant ongoing processes and practices is important for the long-term success of disaster risk reduction and management (Murray et al., 2012). Birkmann and Teichman (2010) caution that while the combination of risk reduction and climate change adaptation strategies may be desirable, different spatial and temporal scales, norm systems, and knowledge types and sources between the two goals can confound their effective combination. Box CC-TC References Birkman, J. and K. von Teichman 2010: Integrating disaster risk reduction and climate change adaptation: key challenges scales, knowledge and norms. Sustainability Science 5: 171-184. Brakenridge, G.R., J.P.M. Syvitski, I. Overeem, S.A. Higgins, A.J. Kettner, J.A. Stewart-Moore, and R. Westerhoff, 2013: Global mapping of storm surges and the assessment of delta vulnerability. Natural Hazards 66: 1295-1312. DOI 10.1007/s11069-012-0317-z IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, 582 pp. Nicholls, R.J., 2007: Adaptation Options for Coastal Areas And Infrastructure: An Analysis For 2030. In Adaptation Options for Coastal Areas And Infrastructure: An Analysis For 2030, 35 pp. Bonn: UNFCCC. Murray V., G. McBean, M. Bhatt, S. Borsch, T.S. Cheong, W.F. Erian, S. Llosa, F. Nadim, M. Nunez, R. Oyun, and A.G. Suarez, 2012: Case studies. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 487-542. Revenga, C., Nackoney, J., Hoshino, E., Kura, Y., Maidens, J., 2003: AS 12 Irrawaddy. Watersheds of the World, Water Resources Institute. Seneviratne, S.I., N. Nicholls, D. Easterling, C.M. Goodess, S. Kanae, J. Kossin, Y. Luo, J. Marengo, K. McInnes, M. Rahimi, M. Reichstein, A. Sorteberg, C. Vera, and X. Zhang, 2012: Changes in climate extremes and their impacts on the natural physical environment. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change (IPCC). Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 109-230. Terry, J., T. F. M. Chui, 2012: Evaluating the fate of freshwater lenses on atoll islands after eustatic sea level rise and cyclone driven inundation: a modelling approach. Global and Planetary Change 88-89, 76-84. References Aarnoudse, E., B. Bluemling, P. Wester, and W. Qu, 2012: The role of collective groundwater institutions in the implementation of direct groundwater regulation measures in Minqin County, China. Hydrogeology Journal, 20 (7), 1213-1221. Acosta-Michlik, L. and V. Espaldon, 2008: Assessing vulnerability of selected farming communities in the Philippines based on a behavioural model of agent's adaptation to global environmental change. Global Environmental Change, 18 (4), 554-563. ADB, 2005: Climate Proofing: A Risk-based Approach to Adaptation. Asian Development Bank, Manila, Philippines, 219 pp. ADB, 2007: Promoting Climate Change Adaptation in Asia and the Pacific. Technical Assistance Report. Project Number: 39343, Asian Development Bank, Manila, Philippines, 1-18 pp. Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 ADB, 2009: The Economics of Climate Change in Southeast Asia: A Regional Review. Asian Development Bank, Manila, Philippines, 255 pp. ADB, 2012: Asian Development Outlook 2012: Confronting Rising Inequality in Asia. Asian Development Outlook, Asian Development Bank, Mandaluyong City, Philippines, 272 pp. ADB, UNESCO, and UNEP, 2012: Green Growth, Resources and Resilience: Environmental Sustainability in Asia and the Pacific. United Nations and Asian Development Bank, Bangkok, Thailand, 134 pp. Adger, W.N., 2003: Social capital, collective action, and adaptation to climate change. Economic Geography, 79 (4), 387-404. Agrawal, A. and N. Perrin, 2008: Climate Adaptation, Local Institutions, and Rural Livelihoods. IFRI Working Paper, W08I-6, International Forestry Resources and Institutions Programme, University of Michigan, Michigan, USA, 17 pp. Ahmed, S. and E. Fajber, 2009: Engendering adaptation to climate variability in Gujarat, India. Gender & Development, 17 (1), 33-50. Aizen, V.B., E.M. Aizen, and V.A. Kuzmichonok, 2007: Glaciers and hydrological changes in the Tien Shan: simulation and prediction. Environmental Research Letters, 2 (4), 10. Aizen, V.B., 2011a: Altai Glaciers. In: Encyclopedia of Snow, Ice and Glaciers [Singh, V.P., P. Singh, and U.K. Haritashya (eds.)]. Springer, Dordrecht, Netherlands, pp. 38-39. Aizen, V.B., 2011b: Pamir Glaciers. In: Encyclopedia of Snow, Ice and Glaciers [Singh, V.P., P. Singh, and U.K. Haritashya (eds.)]. Springer, Dordrecht, Netherlands, pp. 813-815. Aizen, V.B., 2011c: Tien Shan Glaciers. In: Encyclopedia of Snow, Ice and Glaciers [Singh, V.P., P. Singh, and U.K. Haritashya (eds.)]. Springer, Dordrecht, Netherlands, pp. 1179-1181. Al-Bakri, J., A. Suleiman, F. Abdulla, and J. Ayad, 2010: Potential impact of climate change on rainfed agriculture of a semi-arid basin in Jordan. Physics and Chemistry of the Earth, Parts A/B/C, 36 (5 6), 125-134. Alauddin, M. and J. Quiggin, 2008: Agricultural intensification, irrigation and the environment in South Asia: Issues and policy options. Ecological Economics, 65 (1), 111-124. Alcamo, J., N. Dronin, M. Endejan, G. Golubev, and A. Kirilenko, 2007: A new assessment of climate change impacts on food production shortfalls and water availability in Russia. Global Environmental Change, 17 (3 4), 429-444. Alexander, K.S., A. Ryan, and T.G. Measham, 2012: Managed retreat of coastal communities: understanding responses to projected sea level rise. Journal of Environmental Planning and Management, 55 (4), 409-433. Allen, K.M., 2006: Community-based disaster preparedness and climate adaptation: local capacity-building in the Philippines. Disasters, 30 (1), 81-101. Allison, E.H., A.L. Perry, M. Badjeck, W.N. Adger, K. Brown, D. Conway, A.S. Hills, G.M. Pilling, J.D. Reynolds, N.L. Andrew, and N.K. Dulvey, 2009: Vulnerability of national economies to the impacts of climate change on fisheries. Fish and Fisheries, 10, 173-196. Allison, I., 2011: Papua. In: Encyclopedia of Snow, Ice and Glaciers [Singh, V.P., P. Singh, and U.K. Haritashya (eds.)]. Springer, Dordrecht, Netherlands, pp. 815-817. Ananicheva, M.D., M.M. Koreisha, and S. Takahashi, 2005: Assessment of glacier shrinkage from the maximum in the Little Ice Age in the Suntar-Khayata Range, North-East Siberia. Bulletin of Glaciological Research, 22, 9- 17. Ananicheva, M.D., G.A. Kapustin, and M.M. Koreisha, 2006: Glacier changes in the Suntar-Khayata Mountains and Chersky range from the Glacier Inventory of the USSR and satellite images 2001-2003. Data of Glaciological Studies, 101, 163-168. Anderson-Teixeira, K.J., P.K. Snyder, T.E. Twine, S.V. Cuadra, M.H. Costa, and E.H. DeLucia, 2012: Climate- regulation services of natural and agricultural ecoregions of the Americas. Nature Climate Change, 2 (3), 177- 181. Angelsen, A., 2009: Realizing REDD+: National Strategy and Policy Options. Center for International Forestry Research, Bogor, Indonesia, 362 pp. Anisimov, O.A., Y.A. Anokhin, A.N. Krenke, M.D. Ananicheva, P.M. Lurie, and L.T. Myach, 2008: Continental Permafrost and Glaciers. In: Assessment Report on Climate Change and its Consequences in Russian Federation. Volume II Climate Change Consequences [Velichko, A.A. (ed.)]. Planeta Publishing, Moscow, pp. 124-134. Anisimov, O.A., 2009: Stochastic modelling of the active layer thickness under conditions of the current and future climate. Earth's Cryosphere, 13 (3), 36-44. Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Aragao, L.E.O.C., 2012: The rainforest s water pump. Nature, 489, 217-218. Are, F., E. Reimnitz, M. Grigoriev, H.W. Hubberten, and V. Rachold, 2008: The influence of cryogenic processes on the erosional arctic shoreface. Journal of Coastal Research, 24 (1), 110-121. Arias, M.E., T.A. Cochrane, T. Piman, M. Kummu, B.S. Caruso, and T.J. Killeen, 2012: Quantifying changes in flooding and habitats in the Tonle Sap Lake (Cambodia) caused by water infrastructure development and climate change in the Mekong Basin. Journal of Environmental Management, 112, 53-66. Asokan, S.M. and D. Dutta, 2008: Analysis of water resources in the Mahanadi River Basin, India under projected climate conditions. Hydrological Processes, 22 (18), 3589-3603. Atici, C., 2012: Carbon emissions, trade liberalization, and the Japan ASEAN interaction: A group-wise examination. Journal of the Japanese and International Economies, 26, 167-178. Aydinalp, C. and M.S. Cresser, 2008: The effects of global climate change on agriculture. American-Eurasian Journal of Agriculture & Environmental Sciences, 3 (5), 672-676. Ayers, J.M. and S. Huq, 2009: The value of linking mitigation and adaptation: a case study of Bangladesh. Environmental Management, 43 (5), 753-764. Béguin, A., S. Hales, J. Rocklöv, C. Aström, V.R. Louis, and R. Sauerborn, 2011: The opposing effects of climate change and socio-economic development on the global distribution of malaria. Global Environmental Change, 21 (4), 1209-1214. Bagchi, S., 2007: Disease outbreaks in wake of Southeast Asia floods. Canadian Medical Association Journal, 177 (6), 560-560. Bai, F., W. Sang, and J.C. Axmacher, 2011: Forest vegetation responses to climate and environmental change: A case study from Changbai Mountain, NE China. Forest Ecology and Management, 262 (11), 2052-2060. Banu, S., W. Hu, C. Hurst, and S. Tong, 2011: Dengue transmission in the Asia-Pacific region: impact of climate change and socio-environmental factors. Tropical Medicine & International Health, 16 (5), 598-607. Baran, E. and E. Guerin, 2012: Dams, Changes in Sediment Load and Impact on Fish Resources in the Mekong: approach and way forward. Report for the Project A climate resilient mekong: maintaining the flows that nourish life led by the Natural Heritage Institute. WorldFish Center, Phnom Penh, Cambodia, 19 pp. Barange, M. and R.I. Perry, 2009: Physical and ecological impacts of climate change relevant to marine and inland capture fisheries and aquaculture. In: Climate Change Implications for Fisheries and Aquaculture. Overview of Current Scientific Knowledge. FAO Fisheries and Aquaculture Technical Paper 530. [Cochrane, K., C. De Young, D. Soto, and T. Bahri (eds.)]. FAO, Rome, pp. 7-106. Barker, R. and G. Levine, 2012: Water Productivity in Context: The Experiences of Taiwan and the Philippines over the Past Half-Century. IWMI Research Report 145, International Water Management Institute, Colombo, Sri Lanka, 33 pp. Barnett, J.R. and M. Webber, 2010: Accommodating Migration to Promote Adaptation to Climate Change. World Bank Policy Research Working Paper, 62 pp. Barrett, C.B., B.J. Barnett, M.R. Carter, S. Chantarat, J.W. Hansen, A.G. Mude, D.E. Osgood, J.R. Skees, C.G. Turvey, and M.N. Ward, 2007: Poverty Traps and Climate and Weather Risk: Limitations and Opportunities of Index-Based Risk Financing IRI Technical Report 07-03, International Research Institute for Climate and Society, New York, USA, 57 pp. Bates, B.C., Z.W. Kundzewicz, S. Wu, and J.P. Palutikof (eds.), 2008: Climate Change and Water. Technical Paper of the Intergovernmental Panel on Climate Change. IPCC Secretariat, Geneva, 210 pp. Battarbee, R.W., N.J. Anderson, H. Bennion, and G.L. Simpson, 2012: Combining limnological and palaeolimnological data to disentangle the effects of nutrient pollution and climate change on lake ecosystems: problems and potential. Freshwater Biology, 57 (10), 2091-2106. Beaumont, L.J., A. Pitman, S. Perkins, N.E. Zimmermann, N.G. Yoccoz, and W. Thuiller, 2010: Impacts of climate change on the world's most exceptional ecoregions. Proceedings of the National Academy of Sciences of the United States of America, 108 (6), 2306-2311. Begum, R.A., R.D.Z.R.Z. Abidin, and J.J. Pereira, 2011: Initiatives and market mechanisms for climate change actions in Malaysia. Journal of Environmental Science and Technology, 4 (1), 31-40. Bezuijen, M.R., 2011: Wetland Biodiversity & Climate Change Briefing Paper: Rapid Assessment of the Impacts of Climate Change to Wetland Biodiversity in the Lower Mekong Basin. Prepared for the Mekong River Commission by the International Centre for Environmental Management, Hanoi, Vietnam, 37 pp. Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bharati, L., G. Lacombe, P. Gurung, P. Jayakody, C.T. Hoanh, and V. Smakhtin, 2011: The Impacts of Water Infrastructure and Climate Change on the Hydrology of the Upper Ganges River Basin. IWMI Research Report 142, International Water Management Institute, Colombo, Sri Lanka, 36 pp. Bhattachan, A., S. Amatya, T.R. Sedai, S.R. Upreti, and J. Partridge, 2009: Japanese encephalitis in hill and mountain districts, Nepal. Emerging Infectious Diseases, 15 (10), 1691-1692. Bi, J., L. Xu, A. Samanta, Z. Zhu, and R. Myneni, 2013: Divergent Arctic-Boreal vegetation changes between North America and Eurasia over the past 30 years. Remote Sensing, 5 (5), 2093-2112. Bi, P., Y. Zhang, and K.A. Parton, 2007: Weather variables and Japanese encephalitis in the metropolitan area of Jinan city, China. Journal of Infection, 55 (6), 551-556. Bickford, D., S.D. Howard, D.J.J. Ng, and J.A. Sheridan, 2010: Impacts of climate change on the amphibians and reptiles of Southeast Asia. Biodiversity and Conservation, 19 (4), 1043-1062. Biemans, H., I. Haddeland, P. Kabat, F. Ludwig, R.W.A. Hutjes, J. Heinke, W. von Bloh, and D. Gerten, 2011: Impact of reservoirs on river discharge and irrigation water supply during the 20th century. Water Resources Research, 47, 1-15. Birkmann, J., M. Garschagen, F. Kraas, and N. Quang, 2010: Adaptive urban governance: new challenges for the second generation of urban adaptation strategies to climate change. Sustainability Science, 5 (2), 185-206. Biswas, A.K. and K.E. Seetharam, 2008: Achieving water security for Asia. International Journal of Water Resources Development, 24 (1), 145-176. Blanchard, J.L., S. Jennings, R. Holmes, J. Harle, G. Merino, J.I. Allen, J. Holt, N.K. Dulvy, and M. Barange, 2012: Potential consequences of climate change for primary production and fish production in large marine ecosystems. Philosophical Transactions of the Royal Society B-Biological Sciences, 367 (1605), 2979-2989. Blok, D., U. Sass-Klaassen, G. Schaepman-Strub, M.M.P.D. Heijmans, P. Sauren, and F. Berendse, 2011: What are the main climate drivers for shrub growth in Northeastern Siberian tundra? Biogeosciences, 8 (5), 1169-1179. Bogner, J., R. Pipatti, S. Hashimoto, C. Diaz, K. Mareckova, L. Diaz, P. Kjeldsen, S. Monni, A. Faaij, G. Qingxian, Z. Tianzhu, A. Mohammed Abdelrafie, R.T.M. Sutamihardja, and R. Gregory, 2008: Mitigation of global greenhouse gas emissions from waste: conclusions and strategies from the Intergovernmental Panel on Climate Change (IPCC) Fourth Assessment Report. Working Group III (Mitigation). Waste Management & Research, 26 (1), 11-32. Bolch, T., J. Peters, A. Yegorov, B. Pradhan, M. Buchroithner, and V. Blagoveshchensky, 2011: Identification of potentially dangerous glacial lakes in the northern Tien Shan. Natural Hazards, 59 (3), 1691-1714. Bolton, J.J., 2010: The biogeography of kelps (Laminariales, Phaeophyceae): a global analysis with new insights from recent advances in molecular phylogenetics. Helgoland Marine Research, 64 (4), 263-279. Borgaonkar, H.P., A.B. Sikder, and S. Ram, 2011: High altitude forest sensitivity to the recent warming: A tree-ring analysis of conifers from Western Himalaya, India. Quaternary International, 236 (1 2), 158-166. Branch, T.A., B.M. DeJoseph, L.J. Ray, and C.A. Wagner, 2013: Impacts of ocean acidification on marine seafood. Trends in Ecology and Evolution, 28 (3), 178-186. Braun, L.N., W. Hagg, I.V. Severskiy, and G. Young, 2009: Assessment of Snow, Glacier and Water Resources in Asia: Selected Papers from the Workshop in Almaty, Kazakhstan, 2006. Report 8. UNESCO International Hydrological Programme/WMO Hydrology and Water Resource Programme, Koblenz, 186 pp. Brooks, N. and W.N. Adger, 2003: Country Level Risk Measures of Climate-related Natural Disasters and Implications for Adaptation to Climate Change. Tyndall Centre for Climate Change Research, Norwich, UK, 30 pp. Brouwer, R., S. Akter, L. Brander, and E. Haque, 2007: Socioeconomic vulnerability and adaptation to environmental risk: A case study of climate change and flooding in Bangladesh. Risk Analysis, 27 (2), 313-326. Brown, B.E., R.P. Dunne, N. Phongsuwan, and P.J. Somerfield, 2011: Increased sea level promotes coral cover on shallow reef flats in the Andaman Sea, eastern Indian Ocean. Coral Reefs, 30 (4), 867-878. Brown, S.M., 2012: The conservation and development of the Kien Giang Biosphere Reserve: Adaptation to Climate Change. 4 (4), 505-526. Brutsaert, W. and M. Sugita, 2008: Is Mongolia's groundwater increasing or decreasing? The case of the Kherlen River basin. Hydrological Sciences Journal, 53 (6), 1221-1229. Burke, L., K. Reytar, M. Spalding, and A. Perry, 2011: Reefs at Risk Revisited. Washington DC, USA: World Resources Institute. 130 pp. Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Busch, J., R.N. Lubowski, F. Godoy, M. Steininger, A.A. Yusuf, K. Austin, J. Hewson, D. Juhn, M. Farid, and F. Boltz, 2012: Structuring economic incentives to reduce emissions from deforestation within Indonesia. Proceedings of the National Academy of Sciences of the United States of America, 109 (4), 1062-1067. Byg, A. and J. Salick, 2009: Local perspectives on a global phenomenon climate change in Eastern Tibetan villages. Global Environmental Change, 19 (2), 156-166. Cai, W., C. Wang, J. Chen, K. Wang, Y. Zhang, and X. Lu, 2008: Comparison of CO2 emission scenarios and mitigation opportunities in China s five sectors in 2020. Energy Policy, 36, 1181-1194. Cai, X., B.R. Sharma, M.A. Matin, D. Sharma, and S. Gunasinghe, 2010: An Assessment of Crop Water Productivity in the Indus and Ganges River Basins: Current Status and Scope for Improvement. IWMI Research Report 140, International Water Management Institute, Colombo, Sri Lanka, 30 pp. Campbell, B.M., 2009: Beyond Copenhagen: REDD+, agriculture, adaptation strategies and poverty. Global Environmental Change, 19, 397-399. Casassa, G., P. Lopéz, B. Pouyaud, and F. Escobar, 2009: Detection of changes in glacial run-off in alpine basins: examples from North America, the Alps, central Asia and the Andes. Hydrological Processes, 23 (1), 31-41. Champathong, A., D. Komori, M. Kiguchi, T. Sukhapunnaphan, T. Oki, and T. Nakaegawa, 2013: Future projection of mean river discharge climatology for the Chao Phraya River basin. Hydrological Research Letters, 7 (2), 36- 41. Chaturvedi, R.K., R. Gopalakrishnan, M. Jayaraman, G. Bala, N.V. Joshi, R. Sukumar, and N.H. Ravindranath, 2011: Impact of climate change on Indian forests: a dynamic vegetation modeling approach. Mitigation and Adaptation Strategies for Global Change, 16 (2), 119-142. Chen, F., Y.-j. Yuan, W.-s. Wei, Z.-a. Fan, T.-w. Zhang, H.-m. Shang, R.-b. Zhang, S.-l. Yu, C.-r. Ji, and L. Qin, 2012a: Climatic response of ring width and maximum latewood density of Larix sibirica in the Altay Mountains, reveals recent warming trends. Annals of Forest Science, 69 (6), 723-733. Chen, F., Y.-j. Yuan, W.-s. Wei, S.-l. Yu, Z.-a. Fan, R.-b. Zhang, T.-w. Zhang, Q. Li, and H.-m. Shang, 2012b: Temperature reconstruction from tree-ring maximum latewood density of Qinghai spruce in middle Hexi Corridor, China. Theoretical and Applied Climatology, 107 (3-4), 633-643. Chen, F., Y.-j. Yuan, W.-s. Wei, S.-l. Yu, and T.-w. Zhang, 2012c: Tree ring-based winter temperature reconstruction for Changting, Fujian, subtropical region of Southeast China, since 1850: linkages to the Pacific Ocean. Theoretical and Applied Climatology, 109 (1-2), 141-151. Chen, F., Y.-j. Yuan, W.-s. Wei, S.-l. Yu, and T.-w. Zhang, 2012d: Reconstructed temperature for Yong'an, Fujian, Southeast China: Linkages to the Pacific Ocean climate variability. Global and Planetary Change, 86-87, 11-19. Chen, I.C., J.K. Hill, H.J. Shiu, J.D. Holloway, S. Benedick, V.K. Chey, H.S. Barlow, and C.D. Thomas, 2011: Asymmetric boundary shifts of tropical montane Lepidoptera over four decades of climate warming. Global Ecology and Biogeography, 20 (1), 34-45. Chen, J., C.H. Cannon, and H. Hu, 2009: Tropical botanical gardens: at the in situ ecosystem management frontier. Trends in Plant Science, 14 (11), 584-589. Chen, Z., X. Zhang, X. He, N.K. Davi, M. Cui, and J. Peng, 2013: Extension of summer (June August) temperature records for northern Inner Mongolia (1715 2008), China using tree rings. Quaternary International, 283 (0), 21-29. Cheng, G.D. and T.H. Wu, 2007: Responses of permafrost to climate change and their environmental significance, Qinghai-Tibet Plateau. Journal of Geophysical Research-Earth Surface, 112 (F2), 10 pp. Cheung, W.W.L., V.W.Y. Lam, J.L. Sarmiento, K. Kearney, R. Watson, and D. Pauly, 2009: Projecting global marine biodiversity impacts under climate change scenarios. Fish and Fisheries, 10 (3), 235-251. Cheung, W.W.L., V.W.Y. Lam, J.L. Sarmiento, K. Kearney, R. Watson, D. Zeller, and D. Pauly, 2010: Large-scale redistribution of maximum fisheries catch potential in the global ocean under climate change. Global Change Biology, 16 (1), 24-35. Cheung, W.W.L., J.L. Sarmiento, J. Dunne, T.L. Frölicher, V.W.Y. Lam, M.L.D. Palomares, R. Watson, and D. Pauly, 2013: Shrinking of fishes exacerbates impacts of global ocean changes on marine ecosystems. Nature Climate Change, 3 (3), 254-258. Chhatre, A. and A. Agrawal, 2009: Trade-offs and synergies between carbon storage and livelihood benefits from forest communities. Proceedings of the National Academy of Sciences of the United States of America, 106, 17667-17670. Chhokar, K.B., 2010: Higher education and curriculum innovation for sustainable development in India. International Journal of Sustainability in Higher Education, 11 (2), 141-152. Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Choi, S., W.K. Lee, D.A. Kwak, S. Lee, Y. Son, J.H. Lim, and J. Saborowski, 2011: Predicting forest cover changes in future climate using hydrological and thermal indices in South Korea. Climate Research, 49 (3), 229-245. Chou, C., T. Huang, Y. Lee, C. Chen, T. Hsu, and C. Chen, 2011: Diversity of the alpine vegetation in central Taiwan is affected by climate change based on a century of floristic inventories. Botanical Studies, 52 (4), 16 pp. Chou, W.-C., J.-L. Wu, Y.-C. Wang, H. Huang, F.-C. Sung, and C.-Y. Chuang, 2010: Modeling the impact of climate variability on diarrhea-associated diseases in Taiwan (1996 2007). Science of the Total Environment, 409 (1), 43-51. Christensen, J.H., B. Hewitson, A. Busuioc, A. Chen, X. Gao, I. Held, R. Jones, R.K. Kolli, W.-T. Kwon, R. Laprise, V.M. Rueda, L. Mearns, C.G. Menéndez, J. Räisänen, A. Rinke, A. Sarr, and P. Whetton, 2007: Regional Climate Projections, Chapter 11. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor, and H.L. Miller (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 847-940. Chung, J.-Y., Y. Honda, Y.-C. Hong, X.-C. Pan, Y.-L. Guo, and H. Kim, 2009: Ambient temperature and mortality: An international study in four capital cities of East Asia. Science of the Total Environment, 408 (2), 390-396. CIA, 2013: The World Factbook: Country Comparison- Life Expectancy at Birth, Central Intelligence Agency. https://www.cia.gov/library/publications/the-world-factbook/rankorder/2102rank.html [Accessed 27.08.2013]. Clough, Y., J. Barkmann, J. Juhrbandt, M. Kessler, T.C. Wanger, A. Anshary, D. Buchori, D. Cicuzza, K. Darras, D.D. Putra, S. Erasmi, R. Pitopang, C. Schmidt, C.H. Schulze, D. Seidel, I. Steffan-Dewenter, K. Stenchly, S. Vidal, M. Weist, A.C. Wielgoss, and T. Tscharntke, 2011: Combining high biodiversity with high yields in tropical agroforests. Proceedings of the National Academy of Sciences of the United States of America, 108 (20), 8311-8316. Cook, E.R., P.J. Krusic, K.J. Anchukaitis, B.M. Buckley, T. Nakatsuka, and M. Sano, 2012: Tree-ring reconstructed summer temperature anomalies for temperate East Asia since 800 C.E. Climate Dynamics, 1-16. Corlett, R.T., 2011: Impacts of warming on tropical lowland rainforests. Trends in Ecology and Evolution, 26 (11), 606-613. Corlett, R.T. and D.A. Westcott, 2013: Will plant movements keep up with climate change? Trends in Ecology and Evolution, 28 (8), 482-488. Costanza, R.K., I. Paquet, P. King, J. Halimi, S. Sanguanngoi, H., 2011: Planning approaches for water resources development in the Lower Mekong Basin. Portland State University and Mae Fah Luang University, 90 pp. Coulthard, S., 2008: Adapting to environmental change in artisanal fisheries-Insights from a South Indian Lagoon. Global Environmental Change, 18 (3), 479- 489. Cristea, A., D. Hummels, L. Puzzello, and M. Avetisyan, 2013: Trade and the greenhouse gas emissions from international freight transport. Journal of Environmental Economics and Management, 65, 153-173. Crooks, S., D. Herr, J. Tamelander, D. Laffoley, and J. Vandever, 2011: Mitigating Climate Change Through Restoration and Management of Coastal Wetlands and Near-Shore Marine Ecosystems: Challenges and Opportunities. Environment Department Papers 121, World Bank, Washington DC, USA, 64 pp. CWF, GEF, EC, MC, TRF, SCB, and SR, 2009: Shaping Climate-Resilient Development: A Framework for Decision-making, A Report of the Economics of Climate Adaptation Working Group. Climate Works Foundation (CWF), Global Environment Facility (GEF), European Commission (EC), McKinsey & Company (MC), The Rockefeller Foundation (TRF), Standard Chartered Bank (SCB) and Swiss Re (SR), 164 pp. Dahal, S., 2008: Climatic Determinants of Malaria and Kala-Azar in Nepal. Regional Health Forum,32-37. Daniau, A.L., P.J. Bartlein, S.P. Harrison, I.C. Prentice, S. Brewer, P. Friedlingstein, T.I. Harrison-Prentice, J. Inoue, K. Izumi, J.R. Marlon, S. Mooney, M.J. Power, J. Stevenson, W. Tinner, M. Andric, J. Atanassova, H. Behling, M. Black, O. Blarquez, K.J. Brown, C. Carcaillet, E.A. Colhoun, D. Colombaroli, B.A.S. Davis, D. D'Costa, J. Dodson, L. Dupont, Z. Eshetu, D.G. Gavin, A. Genries, S. Haberle, D.J. Hallett, G. Hope, S.P. Horn, T.G. Kassa, F. Katamura, L.M. Kennedy, P. Kershaw, S. Krivonogov, C. Long, D. Magri, E. Marinova, G.M. McKenzie, P.I. Moreno, P. Moss, F.H. Neumann, E. Norström, C. Paitre, D. Rius, N. Roberts, G.S. Robinson, N. Sasaki, L. Scott, H. Takahara, V. Terwilliger, F. Thevenon, R. Turner, V.G. Valsecchi, B. Vanniere, M. Walsh, N. Williams, and Y. Zhang, 2012: Predictability of biomass burning in response to climate changes. Global Biogeochemical Cycles, 26 (4), 11 pp. Dasgupta, S., B. Laplante, C. Meisner, D. Wheeler, and J. Yan, 2009: The impact of sea level rise on developing countries: a comparative analysis. Climatic Change, 93 (3-4), 379-388. Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Davi, N.K., G.C. Jacoby, R.D. D'Arrigo, N. Baatarbileg, J. Li, and A.E. Curtis, 2009: A tree-ring-based drought index reconstruction for far-western Mongolia: 1565-2004. International Journal of Climatology, 29 (10), 1508-1514. Davydov, E.A., G.E. Insarov, and A.K. Sundetpaev, 2013: Lichen monitoring in Katon-Karagai National Park, Eastern Kazakhstan in context of climate change. Problems of Ecological Monitoring and Ecosystem Modelling, 25, 11 pp. Dawe, D., P. Moya, and S. Valencia, 2008: Institutional, policy and farmer responses to drought: El Nino events and rice in the Philippines. Disasters, 33 (2), 291-307. de Jong, R., J. Verbesselt, M.E. Schaepman, and S. de Bruin, 2012: Trend changes in global greening and browning: contribution of short-term trends to longer-term change. Global Change Biology, 18 (2), 642-655. De Silva, S.S. and D. Soto, 2009: Climate change and aquaculture: potential impacts, adaptation and mitigation. In: Climate Change Implications for Fisheries and Aquaculture. Overview of Current Scientific Knowledge. FAO Fisheries and Aquaculture Technical Paper. No. 530 [Cochrane, K., C.D. Young, D. Soto, and T. Bahri (eds.)]. FAO Rome, pp. 151-212. Delbart, N., G. Picard, T. Le Toan, L. Kergoat, S. Quegan, I.A.N. Woodward, D. Dye, and V. Fedotova, 2008: Spring phenology in boreal Eurasia over a nearly century time scale. Global Change Biology, 14 (3), 603-614. Dellink, R., Elzen, M. D., Aiking, H., Bergsma, E., Berkhout, F., Dekker, T. & Gupta, J, 2009: Sharing the Burden of Financing Adaptation to Climate Change. Global Environmental Change 19 (4), 411-421. Delpla, I., A.V. Jung, E. Baures, M. Clement, and O. Thomas, 2009: Impacts of climate change on surface water quality in relation to drinking water production. Environment International, 35 (8), 1225-1233. Dev, S.M., 2011: Climate Change, Rural Livelihoods and Agriculture (Focus on Food Security) in Asia-Pacific Region. Indira Gandhi Institute of Development Research, Mumbai, India, 68 pp. Dev, V. and A. Dash, 2007: Rainfall and malaria transmission in north eastern India. Annals of Tropical Medicine and Parasitology, 101 (5), 457-459. Devi, N.P. and R. Jauhari, 2006: Climatic variables and malaria incidence in Dehradun, Uttaranchal, India. Journal of Vector Borne Diseases, 43 (1), 21. Diaz, S., A. Hector, and D.A. Wardle, 2009: Biodiversity in forest carbon sequestration initiatives: not just a side benefit. Current Opinion in Environmental Sustainability, 1 (1), 55-60. Ding, L., 2012: Application of an operational framework for identifying successful adaptation projects in the Lower Mekong Basin. Asian Journal of Environment and Disaster Risk Management (AJEDM), 4 (4), 379-395. Donato, D.C., J.B. Kauffman, D. Murdiyarso, S. Kurnianto, M. Stidham, and M. Kanninen, 2011: Mangroves among the most carbon-rich forests in the tropics. Nature Geoscience, 4 (5), 293-297. Doney, S.C., M. Ruckelshaus, J.E. Duffy, J.P. Barry, F. Chan, C.A. English, H.M. Galindo, J.M. Grebmeier, A.B. Hollowed, N. Knowlton, J. Polovina, N.N. Rabalais, W.J. Sydeman, and L.D. Talley, 2012: Climate change impacts on marine ecosystems. In: Annual Review of Marine Science, Vol 4 [Carlson, C.A., and S.J. Giovannoni (eds.)], pp. 11-37. Dong, J., G. Zhang, Y. Zhang, and X. Xiao, 2013: Reply to Wang et al.: Snow cover and air temperature affect the rate of changes in spring phenology in the Tibetan Plateau. Proceedings of the National Academy of Sciences of the United States of America, 110 (31), E2856-E2857. Donohue, R.J., M.L. Roderick, T.R. McVicar, and G.D. Farquhar, 2013: Impact of CO2 fertilization on maximum foliage cover across the globe's warm, arid environments. Geophysical Research Letters, 40 (12), 3031-3035. Dorigo, W., R. de Jeu, D. Chung, R. Parinussa, Y. Liu, W. Wagner, and D. Fernández-Prieto, 2012: Evaluating global trends (1988-2010) in harmonized multi-satellite surface soil moisture. Geophysical Research Letters, 39 (18), 7 pp. Douglas, I., 2009: Climate change, flooding and food security in south Asia. Food Security, 1 (2), 127-136. Duan, J., L. Wang, L. Li, and K. Chen, 2010: Temperature variability since A.D. 1837 inferred from tree-ring maximum density of Abies fabri on Gongga Mountain, China. Chinese Science Bulletin, 55 (26), 3015-3022. Dudgeon, D., 2011: Asian river fishes in the Anthropocene: threats and conservation challenges in an era of rapid environmental change. Journal of Fish Biology, 79 (6, Sp. Iss. SI), 1487-1524. Dudgeon, D., 2012: Threats to freshwater biodiversity globally and in the Indo-Burma Biodiversity Hotspot. In: The Status and Distribution of Freshwater Biodiversity in Indo-Burma. [Allen, D.J., K.G. Smith, and W.R.T. Darwall (eds.)]. IUCN, Cambridge, UK, pp. 1-28. Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Dugan, P., A. Delaporte, N. Andrew, M. O'Keefe, and R. Welcomme, 2010: Blue Harvest: Inland Fisheries as an Ecosystem Service. The WorldFish Center Working Papers, United Nations Environment Programme, Penang, Malaysia, 68 pp. Dulamsuren, C., M. Hauck, and C. Leuschner, 2010a: Recent drought stress leads to growth reductions in Larix sibirica in the western Khentey, Mongolia. Global Change Biology, 16 (11), 3024-3035. Dulamsuren, C., M. Hauck, M. Khishigjargal, H.H. Leuschner, and C. Leuschner, 2010b: Diverging climate trends in Mongolian taiga forests influence growth and regeneration of Larix sibirica. Oecologia, 163 (4), 1091-1102. Dulamsuren, C., M. Hauck, H. Leuschner, and C. Leuschner, 2011: Climate response of tree-ring width in Larix sibirica growing in the drought-stressed forest-steppe ecotone of northern Mongolia. Annals of Forest Science, 68 (2), 275-282. Dun, O., 2011: Migration and displacement triggered by floods in the Mekong delta. International Migration, 49 (1), e200-e223. Dutrieux, L.P., H. Bartholomeus, M. Herold, and J. Verbesselt, 2012: Relationships between declining summer sea ice, increasing temperatures and changing vegetation in the Siberian Arctic tundra from MODIS time series (2000 11). Environmental Research Letters, 7 (4), 13 pp. Eichler, A., W. Tinner, S. Brütsch, S. Olivier, T. Papina, and M. Schwikowski, 2011: An ice-core based history of Siberian forest fires since AD 1250. Quaternary Science Reviews, 30 (9-10), 1027-1034. Eliseev, A.V., M.M. Arzhanov, P.F. Demchenko, and Mokhov, II, 2009: Changes in climatic characteristics of Northern Hemisphere extratropical land in the 21st century: Assessments with the IAP RAS climate model. Izvestiya Atmospheric and Oceanic Physics, 45 (3), 271-283. Ensor, J. and R. Berger, 2009: Understanding Climate Change Adaptation: Lessons from Community-Based Approaches. Practical Action, UK. Epstein, H.E., M.K. Raynolds, D.A. Walker, U.S. Bhatt, C.J. Tucker, and J.E. Pinzon, 2012: Dynamics of aboveground phytomass of the circumpolar Arctic tundra during the past three decades. Environmental Research Letters, 7 (1), 13 pp. Estudillo, J.P. and K. Otsuka, 2010: Rural poverty and income dynamics in Southeast Asia. Handbook of Agricultural Economics, 4, 3434-3468. Evans, A., 2010: Resource Scarcity, Climate Change and the Risk of Violent Conflict. Background Paper, World Development Report 2011, 23 pp. FAO, 2010: State of World Fisheries and Agriculture 2010. FAO Fisheries and Aquaculture Department, Rome, 197 pp. FAOSTAT, 2011: Faostat Popstat. http://faostat.fao.org/faostat [Accessed 06.29]. Fargione, J.E., R.J. Plevin, and J.D. Hill, 2010: The ecological impact of biofuels. In: Annual Review of Ecology, Evolution, and Systematics, Vol 41 [Futuyma, D.J., H.B. Shafer, and D. Simberloff (eds.)], pp. 351-377. Flannigan, M.D., M.A. Krawchuk, W.J. de Groot, B.M. Wotton, and L.M. Gowman, 2009: Implications of changing climate for global wildland fire. International Journal of Wildland Fire, 18 (5), 483-507. Fleischer, A., R. Mendelsohn, and A. Dinar, 2011: Building Agricultural Technologies to Adapt to Climate Change. Technological Forecasting and Social Change, 75, 982-990. Forrest, J.L., E. Wikramanayake, R. Shrestha, G. Areendran, K. Gyeltshen, A. Maheshwari, S. Mazumdar, R. Naidoo, G.J. Thapa, and K. Thapa, 2012: Conservation and climate change: Assessing the vulnerability of snow leopard habitat to treeline shift in the Himalaya. Biological Conservation, 150 (1), 129-135. Fuchs, R., M. Conran, and E. Louis, 2011: Climate change and Asia's coastal urban cities: Can they meet the challenge? Environment and Urbanization Asia, 2 (1), 13-28. Fung, F., A. Lopez, and M. New, 2011: Water availability in +2 degrees C and +4 degrees C worlds. Philosophical Transactions of the Royal Society A - Mathematical Physical and Engineering Sciences, 369 (1934), 99-116. Gaillard, J.-C., C.C. Liamzon, and J.D. Villanueva, 2007: Natural disaster? A retrospect into the causes of the late- 2004 typhoon disaster in Eastern Luzon, Philippines. Environmental Hazards, 7 (4), 257-270. Game, E.T., G. Lipsett-Moore, E. Saxon, N. Peterson, and S. Sheppard, 2011: Incorporating climate change adaptation into national conservation assessments. Global Change Biology, 17 (10), 3150-3160. García-López, J.M. and C. Allué, 2013: Modelling future no-analogue climate distributions: A world-wide phytoclimatic niche-based survey. Global and Planetary Change, 101 (0), 1-11. Gardner, A.S., G. Moholdt, J.G. Cogley, B. Wouters, A.A. Arendt, J. Wahr, E. Berthier, R. Hock, W.T. Pfeffer, G. Kaser, S.R.M. Ligtenberg, T. Bolch, M.J. Sharp, J.O. Hagen, M.R. van den Broeke, and F. Paul, 2013: A reconciled estimate of glacier contributions to sea level rise: 2003 to 2009. Science, 340 (6134), 852-857. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Garg, A., R. Dhiman, S. Bhattacharya, and P. Shukla, 2009: Development, malaria and adaptation to climate change: a case study from India. Environmental Management, 43 (5), 779-789. Garschagen, M., F.G. Renaud, and J. Birkmann, 2011: Dynamic Resilience of Peri-Urban Agriculturalists in the Mekong Delta Under Pressures of Socio-Economic Transformation and Climate Change. In: Environmental Change and Agricultural Sustainability in the Mekong Delta [Stewart, M.A., and P.A. Coclanis (eds.)]. Springer Netherlands, Dordrecht, pp. 141-163. Garschagen, M. and F. Kraas, 2011: Urban Climate Change Adaptation in the Context of Transformation: Lessons from Vietnam. In: Resilient Cities: Cities and Adaptation to Climate Change - Proceedings of the Global Forum 2010 [Otto-Zimmermann, K. (ed.)], pp. 131-139. Gautam, P.K., 2012: Climate change and conflict in South Asia. Strategic Analysis, 36 (1), 32-40. Gautam, R., N.C. Hsu, T.F. Eck, B.N. Holben, S. Janjai, T. Jantarach, S.-C. Tsay, and W.K. Lau, 2013: Characterization of Aerosols over the Indochina Peninsula from Satellite-Surface Observations During Biomass Burning Pre-Monsoon Season. Atmospheric Environment, 78, 51-59. GAVI, 2011: GAVI Alliance Progress Report 2011. Access at www.gavialliance.org, 68 pp. Ge, Q., H. Wang, and J. Dai, 2013: Simulating changes in the leaf unfolding time of 20 plant species in China over the twenty-first century. International Journal of Biometeorology. Gedan, K.B., M.L. Kirwan, E. Wolanski, E.B. Barbier, and B.R. Silliman, 2011: The present and future role of coastal wetland vegetation in protecting shorelines: answering recent challenges to the paradigm. Climatic Change, 106 (1), 7-29. Gessner, U., V. Naeimi, I. Klein, C. Kuenzer, D. Klein, and S. Dech, 2013: The relationship between precipitation anomalies and satellite-derived vegetation activity in Central Asia. Global and Planetary Change, 2, 14 pp. Gigli, S. and S. Agrawala, 2007: Stocktaking of Progress on Integrating Adaptation to Climate Change into Development Co-operation Activities. OECD, Paris, 83 pp. Gilman, E.L., J. Ellison, N.C. Duke, and C. Field, 2008: Threats to mangroves from climate change and adaptation options: A review. Aquatic Botany, 89 (2), 237-250. Giri, C., E. Ochieng, L.L. Tieszen, Z. Zhu, A. Singh, T. Loveland, J. Masek, and N. Duke, 2011: Status and distribution of mangrove forests of the world using earth observation satellite data. Global Ecology and Biogeography, 20 (1), 154-159. Glazyrin, G.E. and U.U. Tadzhibaeva, 2011: Climate change in the high mountains of Central Asia in the late XX century. Lëd i Sneg, 114, 12-15 [In Russian]. Goetz, S., H. Epstein, U. Bhatt, G. Jia, J. Kaplan, H. Lischke, Q. Yu, A. Bunn, A. Lloyd, D. Alcaraz-Segura, P.A. Beck, J. Comiso, M. Raynolds, and D. Walker, 2011: Recent changes in Arctic vegetation: Satellite observations and simulation model predictions. In: Eurasian Arctic Land Cover and Land Use in a Changing Climate [Gutman, G., and A. Reissell (eds.)]. Springer Netherlands, pp. 9-36. Golubyatnikov, L.L. and E.A. Denisenko, 2007: Model estimates of climate change impact on habitats of zonal vegetation for the plain territories of Russia. Biology Bulletin, 34 (2), 170-184. Green, E.P. and F.T. Short, 2003: World Atlas of Seagrasses. UNEP-WCMC, University of California Press, Berkeley, Los Angeles and London, 298 pp. Griffin, D.W., 2007: Atmospheric movement of microorganisms in clouds of desert dust and implications for human health. Clinical Microbiology Reviews, 20 (3), 459-477. Grigor ev, A.A., P.A. Moiseev, and Z.Y. Nagimov, 2013: Dynamics of the timberline in high mountain areas of the Nether-Polar Urals under the influence of current climate change. Russian Journal of Ecology, 44 (4), 312-323. Grumbine, R.E., J. Dore, and J. Xu, 2012: Mekong hydropower: drivers of change and governance challenges. Frontiers in Ecology and the Environment, 10 (2), 91-98. Guan, D. and K. Hubacek, 2007: Assessment of regional trade and virtual water flows in China. Ecological Economics, 61 (1), 159-170. Guan, P., D. Huang, M. He, T. Shen, J. Guo, and B. Zhou, 2009: Investigating the effects of climatic variables and reservoir on the incidence of hemorrhagic fever with renal syndrome in Huludao City, China: a 17-year data analysis based on structure equation model. BMC Infectious Diseases, 9 (1), 109. Guariguata, M.R., J.P. Cornelius, B. Locatelli, C. Forner, and G.A. Sánchez-Azofeifa, 2008: Mitigation needs adaptation: Tropical forestry and climate change. Mitigation and Adaptation Strategies for Global Change, 13 (8), 793-808. Gumilang, H., K. Mukhopadhyay, and P.J. Thomassin, 2011: Economic and environmental impacts of trade liberalization: The case of Indonesia. Economic Modelling, 28 (3), 1030-1041. Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Guo, Y., J. Wang, G. Li, Y. Zheng, W. He, and X. Pan, 2009: Association between ambient temperature and hospital emergency room visits for cardiovascular diseases: a case-crossover study. Chinese Journal of Epidemiology, 30 (8), 810. Hadano, M., K.N. Nasahara, T. Motohka, H.M. Noda, K. Murakami, and M. Hosaka, 2013: High-resolution prediction of leaf onset date in Japan in the 21st century under the IPCC A1B scenario. Ecology and Evolution, 3 (6), 1798-1807. Haggblade, S., P. Hazell, and T. Reardon, 2009: Transforming the Rural Nonfarm Economy: Opportunities and Threats in the Developing World International Food Policy Research Institute, Washington DC, USA, 58 pp. Haggblade, S., P. Hazell, and T. Reardon, 2010: The rural non-farm economy: Prospects for growth and poverty reduction. World Development, 38 (10), 1429-1441. Haghdoost, A.A., N. Alexander, and J. Cox, 2008: Modelling of malaria temporal variations in Iran. Tropical Medicine & International Health, 13 (12), 1501-1508. Haines, A., A.J. McMichael, K.R. Smith, I. Roberts, J. Woodcock, A. Markandya, B.G. Armstrong, D. Campbell- Lendrum, A.D. Dangour, M. Davies, N. Bruce, C. Tonne, M. Barrett, and P. Wilkinson, 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: overview and implications for policy makers. Lancet, 374 (9707), 2104-2114. Hallegatte, S. and J. Corfee-Morlot, 2011: Understanding climate change impacts, vulnerability and adaptation at city scale: an introduction. Climatic Change, 104 (1), 1-12. Halls, A.S., 2009: Addressing fisheries in the Climate Change and Adaptation Initiative Fisheries and climate change. Catch and Culture: Fisheries Research and Development in the Mekong Region, 15 (1), 48 pp. Hamilton, S.K., 2010: Biogeochemical implications of climate change for tropical rivers and floodplains. Hydrobiologia, 657 (1), 19-35. Handmer, J., Y. Honda, Z.W. Kundzewicz, N. Arnell, G. Benito, H. J., I.F. Mohamed, P. Peduzzi, S. Wu, B. Sherstyukov, K. Takahashi, and Z. Yan, 2012: Changes in impacts of climate extremes: human systems and ecosystems. In: Managing the Risks of Extreme Events and Disasters to Advanced Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, P. G.K., S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, UK, pp. 231-290. Hannah, L., 2010: A global conservation system for climate-change adaptation. Conservation Biology, 24 (1), 70-77. Hanson, S., R. Nicholls, N. Ranger, S. Hallegatte, J. Corfee-Morlot, C. Herweijer, and J. Chateau, 2011: A global ranking of port cities with high exposure to climate extremes. Climatic Change, 104 (1), 89-111. Harris, A.M., F. Chowdhury, Y.A. Begum, A.I. Khan, A.S. Faruque, A.-M. Svennerholm, J.B. Harris, E.T. Ryan, A. Cravioto, and S.B. Calderwood, 2008: Shifting prevalence of major diarrheal pathogens in patients seeking hospital care during floods in 1998, 2004, and 2007 in Dhaka, Bangladesh. American Journal of Tropical Medicine and Hygiene, 79 (5), 708-714. Harshita, T., 2013: Impact of drought on rice based agriculture economy- A case study of livelihood security in rural areas of Sultanpur District of Uttar Pradesh. Voice of Intellectual Man-An International Journal, 3 (1), 137-148. Hashizume, M., B. Armstrong, S. Hajat, Y. Wagatsuma, A.S. Faruque, T. Hayashi, and D.A. Sack, 2007: Association between climate variability and hospital visits for non-cholera diarrhoea in Bangladesh: effects and vulnerable groups. International Journal of Epidemiology, 36 (5), 1030-1037. Hashizume, M., Y. Wagatsuma, A.S.G. Faruque, T. Hayashi, P.R. Hunter, B. Armstrong, and D.A. Sack, 2008: Factors determining vulnerability to diarrhoea during and after severe floods in Bangladesh. Journal of Water and Health, 6 (3), 323-332. Hashizume, M., Y. Wagatsuma, T. Hayashi, S.K. Saha, K. Streatfield, and M. Yunus, 2009: The effect of temperature on mortality in rural Bangladesh a population-based time-series study. International Journal of Epidemiology, 38 (6), 1689-1697. Hashizume, M., A.S. Faruque, Y. Wagatsuma, T. Hayashi, and B. Armstrong, 2010: Cholera in Bangladesh: climatic components of seasonal variation. Epidemiology, 21 (5), 706-710. Hashizume, M., A. Faruque, T. Terao, M. Yunus, K. Streatfield, T. Yamamoto, and K. Moji, 2011: The Indian Ocean dipole and cholera incidence in Bangladesh: a time-series analysis. Environmental Health Perspectives, 119 (2), 239. Hashizume, M., A.M. Dewan, T. Sunahara, M.Z. Rahman, and T. Yamamoto, 2012: Hydroclimatological variability and dengue transmission in Dhaka, Bangladesh: a time-series study. BMC Infectious Diseases, 12 (1). Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Heltberg, R., R. Prabhu, and H. Gitay, 2010: Community-based Adaptation: Lessons from the Development Marketplace 2009 on Adaptation to Climate Change. Fondazione Eni Enrico Mattei Working Papers, Paper 473., Fondazione Eni Enrico Mattei, Milan, Italy, 64 pp. Hertel, T.W., M.B. Burke, and D.B. Lobell, 2010: The poverty implications of climate-induced crop yield changes by 2030. Global Environmental Change, 20 (4), 577-585. Higgins, S.I. and S. Scheiter, 2012: Atmospheric CO2 forces abrupt vegetation shifts locally, but not globally. Nature, 488 (7410), 209-212. Hii, Y.L., J. Rocklöv, N. Ng, C.S. Tang, F.Y. Pang, and R. Sauerborn, 2009: Climate variability and increase in intensity and magnitude of dengue incidence in Singapore. Global Health Action, 2, 1 pp. Hoegh-Guldberg, O., 2011: Coral reef ecosystems and anthropogenic climate change. Regional Environmental Change, 11 (Suppl 1) S215-S227. Honda, Y. and M. Ono, 2009: Issues in health risk assessment of current and future heat extremes. Global Health Action, 2. Hsieh, Y.H. and C. Chen, 2009: Turning points, reproduction number, and impact of climatological events for multi wave dengue outbreaks. Tropical Medicine & International Health, 14 (6), 628-638. Huang, D., G. Peng, G. Junqiao, W. Ping, and Z. Baosen, 2008: Investigating the effects of climate variations on bacillary dysentery incidence in northeast China using ridge regression and hierarchical cluster analysis. BMC Infectious Diseases, 8. Huang, W., H. Kan, and S. Kovats, 2010: The impact of the 2003 heat wave on mortality in Shanghai, China. Science of the Total Environment, 408 (11), 2418-2420. Huang, X.-X., T.-J. Wang, F. Jiang, J.-B. Liao, Y.-F. Cai, C.-Q. Yin, J.-L. Zhu, and Y. Han, 2013: Studies on a severe dust storm in East Asia and its impact on the air quality of Nanjing, China. Aerosol and Air Quality Research, 13 (1), 179-193. Hughes, A.C., C. Satasook, P.J.J. Bates, S. Bumrungsri, and G. Jones, 2012: The projected effects of climatic and vegetation changes on the distribution and diversity of Southeast Asian bats. Global Change Biology, 18 (6), 1854-1865. Huigen, M.G.A. and I.C. Jens, 2006: Socio-economic impact of super typhoon Harurot in San Mariano, Isabela, the Philippines. World Development, 34 (12), 2116-2136. Huq, A., R.B. Sack, A. Nizam, I.M. Longini, G.B. Nair, A. Ali, J.G. Morris, M.H. Khan, A.K. Siddique, and M. Yunus, 2005: Critical factors influencing the occurrence of Vibrio cholerae in the environment of Bangladesh. Applied and Environmental Microbiology, 71 (8), 4645-4654. Huq, S.R., H., 2007: Community Based Adaptation: A Vital Approach to the Threat Climate Change Poses to the Poor. IIED Briefing, International Institute for Environment and Development, UK, 2 pp. Husain, T. and J.R. Chaudhary, 2008: Human health risk assessment due to global warming a case study of the Gulf countries. International Journal of Environmental Research and Public Health, 5 (4), 204-212. Hussain, S.S. and M. Mudasser, 2007: Prospects for wheat production under changing climate in mountain areas of Pakistan An econometric analysis. Agricultural Systems, 94 (2), 494-501. Hyatt, O.M., B. Lemke, and T. Kjellstrom, 2010: Regional maps of occupational heat exposure: past, present, and potential future. Global Health Action, 3 (0). ICEM, 2010: Climate Change Baseline Assessment Working Paper. MRC Strategic Environmental Assessment (SEA) of Hydropower on the Mekong Mainstream: Baseline Report, International Centre for Environmental Management, Hanoi, Viet Nam, 24 pp. IFAD, 2010: Rural Poverty Report 2011: New Reality, New Challenges, New Opportunities for Tomorrow s Generation. International Fund for Agricultural Development, Rome, Italy, 317 pp. Im, E.S., I.W. Jung, H. Chang, D.H. Bae, and W.T. Kwon, 2010: Hydroclimatological response to dynamically downscaled climate change simulations for Korean basins. Climatic Change, 100 (3-4), 485-508. Immerzeel, W.W., L.P.H. Van Beek, and M.F.P. Bierkens, 2010: Climate change will affect the Asian water towers. Science, 328 (5984), 1382-1385. Insarov, G.E., O.K. Borisova, M.D. Korzukhin, V.N. Kudeyarov, A.A. Minin, A.V. Olchev, S.M. Semenov, A.A. Sirin, and V.I. Kharuk, 2012: Chapter 6: Terrestrial Ecosystems. In: Methods for Assessment of Climate Change Impacts on Physical and Biological Systems [Semenov, S.M. (ed.)]. Planet Publishing, Moscow, pp. 190-265. IPCC, 2007: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group I to the Fourth Assessment Report of the International Panel on the Climate Change. [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J.v.d. Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, p. 976. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change. [Field, C.B., V. Barros, T.F. Stocker, Q. Dahe, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, UK, and New York, USA, p. 582. Iqbal, M.M., M.A. Goheer, and A.M. Khan, 2009: Climate change aspersions on food security of Pakistan. Science Vision, 15 (No.1), 15-23. IRG, 2010: USAID Asia-Pacific Regional Climate Change Needs Assessment Final Report: Findings and Recommendations. International Resources Group, Washington D.C., 145 pp. Ishizuka, W. and S. Goto, 2012: Modeling intraspecific adaptation of Abies sachalinensis to local altitude and responses to global warming, based on a 36-year reciprocal transplant experiment. Evolutionary Applications, 5 (3), 229-244. Islam, M.K., Merlo, J., Kawachi, I., Lindström, M. and Gerdtham, U.-G, 2006: Social Capital and Health: Does Egalitarianism Matter? A Literature Review. International Journal for Equity in Health, 5 (1), 3. IUCN, 2009: Ecosystem-based Adaptation: A Natural Response to Climate Change., IUCN, Switzerland, 20 pp. Iwasaki, S., B.H.N. Razafindrabe, and R. Shaw, 2009: Fishery livelihoods and adaptation to climate change: a case study of Chilika lagoon, India. Mitigation and Adaptation Strategies for Global Change, 14 (4), 339-355. Janvry, A. and E. Sadoulet, 2010: Agricultural growth and poverty reduction: Additional evidence. The World Bank Research Observer, 25 (1), 1-20. Jarvis, A., C. Lau, S. Cook, E. Wollenberg, J. Hansen, O. Bonilla, and A. Challinor, 2011: An integrated adaptation and mitigation framework for developing agricultural research: Synergies and tradeoffs. Experimental Agriculture, 47, 185-203. Jasparro, C. and J. Taylor, 2008: Climate change and regional vulnerability to transnational security threats in Southeast Asia. Geopolitics, 13 (2), 232-256. Jeong, S.-J., C.-H. Ho, B.-M. Kim, S. Feng, and D. Medvigy, 2013: Non-linear response of vegetation to coherent warming over northern high latitudes. Remote Sensing Letters, 4 (2), 123-130. Jevanandam, N., A.G.R. Goh, and R.T. Corlett, 2013: Climate warming and the potential extinction of fig wasps, the obligate pollinators of figs. Biology Letters, 9 (3), 5 pp. Jian, J., P.J. Webster, and C.D. Hoyos, 2009: Large-scale controls on Ganges and Brahmaputra river discharge on intraseasonal and seasonal time-scales. Quarterly Journal of the Royal Meteorological Society, 135 (639), 353- 370. Jin, Z., Q. Zhuang, J.-S. He, T. Luo, and Y. Shi, 2013: Phenology shift from 1989 to 2008 on the Tibetan Plateau: an analysis with a process-based soil physical model and remote sensing data. Climatic Change, 119 (2), 435-449. Johnston, F.H., S.B. Henderson, Y. Chen, J.T. Randerson, M. Marlier, R.S. DeFries, P. Kinney, D.M. Bowman, and M. Brauer, 2012: Estimated global mortality attributable to smoke from landscape fires. Environmental Health Perspectives, 120 (5), 695-701. Jones, H.P.H., D.G. Zavaleta, E.S., 2012: Harnessing nature to help people adapt to climate change. Nature Climate Change, 2 504-509. Jones, L. and E. Boyd, 2011: Exploring social barriers to adaptation: insights from Western Nepal. Global Environmental Change 21 (1), 1262-1274. Joubert, D., J. Thomsen, and O. Harrison, 2011: Safety in the heat: A comprehensive program for prevention of heat illness among workers in Abu Dhabi, United Arab Emirates. American Journal of Public Health, 101 (3), 395- 398. Jump, A.S., T.J. Huang, and C.H. Chou, 2012: Rapid altitudinal migration of mountain plants in Taiwan and its implications for high altitude biodiversity. Ecography, 35 (3), 204-210. Kan, H., S.J. London, H. Chen, G. Song, G. Chen, L. Jiang, N. Zhao, Y. Zhang, and B. Chen, 2007: Diurnal temperature range and daily mortality in Shanghai, China. Environmental Research, 103 (3), 424-431. Kan, H., R. Chen, and S. Tong, 2012: Ambient air pollution, climate change, and population health in China. Environment International, 42, 10-19. Kang, S., B. Yang, and C. Qin, 2012: Recent tree-growth reduction in north central China as a combined result of a weakened monsoon and atmospheric oscillations. Climatic Change, 115 (3-4), 519-536. Kaplan, J.O. and M. New, 2006: Arctic climate change with a 2 degrees C global warming: Timing, climate patterns and vegetation change. Climatic Change, 79 (3-4), 213-241. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Karim, M.F. and N. Mimura, 2008: Impacts of climate change and sea-level rise on cyclonic storm surge floods in Bangladesh. Global Environmental Change, 18 (3), 490-500. Kariyeva, J., W.D. Leeuwen, and C. Woodhouse, 2012: Impacts of climate gradients on the vegetation phenology of major land use types in Central Asia (1981 2008). Frontiers of Earth Science, 6 (2), 206-225. Kawaguchi, L., B. Sengkeopraseuth, R. Tsuyuoka, N. Koizumi, H. Akashi, P. Vongphrachanh, H. Watanabe, and A. Aoyama, 2008: Seroprevalence of leptospirosis and risk factor analysis in flood-prone rural areas in Lao PDR. American Journal of Tropical Medicine and Hygiene, 78 (6), 957-961. Kazama, S., T. Aizawa, T. Watanabe, P. Ranjan, L. Gunawardhana, and A. Amano, 2012: A quantitative risk assessment of waterborne infectious disease in the inundation area of a tropical monsoon region. Sustainability Science, 7 (1), 45-54. Kelkar, U., K.K. Narula, V.P. Sharma, and U. Chandna, 2008: Vulnerability and adaptation to climate variability and water stress in Uttarakhand State, India. Global Environmental Change, 18 (4), 564-574. Keskinen, M.C., S.Kummu, M. Nuorteva, P. Snidvongs, A.Varis, and O. Vastila, K., 2010: Climate change and water resources in the lower Mekong river basin: Putting adaptation into context. Journal of Water and Climate Change, 1 (2), 103-117. Kharuk, V.I., K.J. Ranson, S.T. Im, and M.M. Naurzbaev, 2006: Forest-tundra larch forests and climatic trends. Russian Journal of Ecology, 37 (5), 291-298. Kharuk, V.I., S.T. Im, and M.L. Dvinskaya, 2010a: Forest-tundra ecotone response to climate change in the Western Sayan Mountains, Siberia. Scandinavian Journal of Forest Research, 25 (3), 224-233. Kharuk, V.I., S.T. Im, M.L. Dvinskaya, and K.J. Ranson, 2010b: Climate-induced mountain tree-line evolution in southern Siberia. Scandinavian Journal of Forest Research, 25 (5), 446-454. Kharuk, V.I., K.J. Ranson, and M.L. Dvinskaya, 2010c: Evidence of evergreen conifers invasion into larch dominated forests during recent decades. In: Environmental Change in Siberia: Earth Observation, Field Studies and Modelling [Balzter, H. (ed.)]. Springer, Dordrecht, pp. 53-65. Kharuk, V.I., K.J. Ranson, M.L. Dyinskaya, and S.T. Im, 2010d: Siberian pine and larch response to climate warming in the southern Siberian mountain forest: tundra ecotone. In: Environmental Change in Siberia: Earth Observation, Field Studies and Modelling [Balzter, H. (ed.)]. Springer, Dordrecht, pp. 115-132. Kharuk, V.I., K.J. Ranson, S.T. Im, and A.S. Vdovin, 2010e: Spatial distribution and temporal dynamics of high- elevation forest stands in southern Siberia. Global Ecology and Biogeography, 19 (6), 822-830. Kharuk, V.I., K.J. Ranson, P.A. Oskorbin, S.T. Im, and M.L. Dvinskaya, 2013: Climate induced birch mortality in Trans-Baikal lake region, Siberia. Forest Ecology and Management, 289, 385-392. Kheir, S., A. Alahmed, M. Al Kuriji, and S.F. Al Zubyani, 2010: Distribution and seasonal activity of mosquitoes in al Madinah Al Munwwrah, Saudi Arabia. Journal of the Egyptian Society of Parasitology, 40 (1), 215. Khim, L., 2012: Climate resilience in rural Cambodia: Adaptation mainstreaming, water resource management and agricultural practice. Asian Journal of Environment and Disaster Risk Management (AJEDM) 2012 Special Issue, 4 (4), 463-484. Kim, B.S., H.S. Kim, B.H. Seoh, and N.W. Kim, 2007: Impact of climate change on water resources in Yongdam Dam Basin, Korea. Stochastic Environmental Research and Risk Assessment, 21 (4), 355-373. Kim, H., J.-S. Ha, and J. Park, 2006: High temperature, heat index, and mortality in 6 major cities in South Korea. Archives of Environmental & Occupational Health, 61 (6), 265-270. Kim Oanh, N.T. and K. Leelasakultum, 2011: Analysis of meteorology and emission in haze episode prevalence over mountain-bounded region for early warning. Science of the Total Environment, 409 (11), 2261-2271. Kim, S.-H. and J.-Y. Jang, 2010: Correlations between climate change-related infectious diseases and meteorological factors in Korea. Journal of Preventive Medicine and Public Health, 43 (5), 436-444. Kintisch, E., 2013: Can coastal marshes rise above it all? Science, 341 (6145), 480-481. Klein, R.J.T., S. Huq, F. Denton, T.E. Downing, R.G. Richels, J.B. Robinson, and F.L. Toth, 2007: Inter- relationships between adaptation and mitigation. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J.v.d. Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, pp. 745-777. Klorvuttimontara, S., C.J. McClean, and J.K. Hill, 2011: Evaluating the effectiveness of Protected Areas for conserving tropical forest butterflies of Thailand. Biological Conservation, 144 (10), 2534-2540. Knox, J., T. Hess, A. Daccache, and T. Wheeler, 2012: Climate change impacts on crop productivity in Africa and South Asia. Environmental Research Letters, 7 (3), 034032. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Knox, J.W., T.M. Hess, A. Daccache, and M.P. Ortola, 2011: What are the Projected Impacts of Climate Change on Food Crop Productivity in Africa and South Asia? DFID Systematic Review, Canfield University, Canfield, UK, 77 pp. Knutson, T.R., J.L. McBride, J. Chan, K. Emanuel, G. Holland, C. Landsea, I. Held, J.P. Kossin, A.K. Srivastava, and M. Sugi, 2010: Tropical cyclones and climate change. Nature Geoscience, 3 (3), 157-163. Ko, C.Y., T.L. Root, S.H. Lin, S.H. Schneider, and P.F. Lee, 2012: Global change projections for Taiwan island birds: linking current and future distributions. Nature Conservation, 2, 21-40. Kolmannskog, V.O., 2008: Future Floods of Refugees: A Comment on Climate Change, Conflict and Forced Migration. Norwegian Refugee Council, Olso, Norway, 44 pp. Konovalov, V. and L. Desinov, 2007: Remote sensing monitoring of the long-term regime of the Pamirs glaciers. International Association of Hydrological Sciences Publications, 316, 149-156. Korzukhin, M.D. and Y.L. Tcelniker, 2010: Model analysis of present ranges for forest tree species in Russia and their changes under two climatic scenarios. Problems of Ecological Monitoring and Ecosystem Modelling, 23, 249-268. Kostianoy, A.G. and A.N. Kosarev, 2010: The Aral Sea Environment. Handbook of Environmental Chemistry. Springer, Berlin and Heidelberg, 1st ed., 335 pp. Kranz, N., T. Menniken, and J. Hinkel, 2010: Climate change adaptation strategies in the Mekong and Orange- Senqu basins: What determines the state-of-play? Environmental Science & Policy, 13 (7), 648-659. Kroeker, K.J., R.L. Kordas, R. Crim, I.E. Hendriks, L. Ramajo, G.S. Singh, C.M. Duarte, and J.-P. Gattuso, 2013: Impacts of ocean acidification on marine organisms: quantifying sensitivities and interaction with warming. Global Change Biology, 19 (6), 1884-1896. Kukavskaya, E.A., A.J. Soja, A.P. Petkov, E.I. Ponomarev, G.A. Ivanova, and S.G. Conard, 2013: Fire emissions estimates in Siberia: evaluation of uncertainties in area burned, land cover, and fuel consumption. Canadian Journal of Forest Research, 43 (5), 493-506. Kumagai, T. and A. Porporato, 2012: Drought-induced mortality of a Bornean tropical rain forest amplified by climate change. Journal of Geophysical Research-Biogeosciences, 117 (G2), 13 pp. Kumar, R.H., K. Venkaiah, N. Arlappa, S. Kumar, G. Brahmam, and K. Vijayaraghavan, 2005: Diet and Nutritional Situation of the Population in the Severely Drought Affected Areas of Gujarat. Journal of Human Ecology, 18 (4), 319-326. Kumpula, T., A. Pajunen, E. Kaarlejarvi, B.C. Forbes, and F. Stammler, 2011: Land use and land cover change in Arctic Russia: Ecological and social implications of industrial development. Global Environmental Change, 21, 550-562. La Sorte, F.A. and W. Jetz, 2010: Projected range contractions of montane biodiversity under global warming. Proceedings of the Royal Society B-Biological Sciences, 277 (1699), 3401-3410. Laczko, F. and C. Aghazarm, 2009: Migration, Environment and Climate Change: Assessing the Evidence. International Organization for Migration, 448 pp. Lal, M., 2011: Implications of climate change in sustained agricultural productivity in South Asia. Regional Environmental Change, 11, S79-S94. Laneri, K., A. Bhadra, E.L. Ionides, M. Bouma, R.C. Dhiman, R.S. Yadav, and M. Pascual, 2010: Forcing Versus Feedback: Epidemic Malaria and Monsoon Rains in Northwest India. PLoS Computational Biology, 6 (9), e1000898. Langkulsen, U., N. Vichit-Vadakan, and S. Taptagaporn, 2010: Health impact of climate change on occupational health and productivity in Thailand. Global Health Action, 3. Lantuit, H., P.P. Overduin, N. Couture, S. Wetterich, F. Aré, D. Atkinson, J. Brown, G. Cherkashov, D. Drozdov, D.L. Forbes, A. Graves-Gaylord, M. Grigoriev, H.-W. Hubberten, J. Jordan, T. Jorgenson, R.S. Odegard, S. Ogorodov, W.H. Pollard, V. Rachold, S. Sedenko, S. Solomon, F. Steenhuisen, I. Streletskaya, and A. Vasiliev, 2012: The Arctic Coastal Dynamics Database: A New Classification Scheme and Statistics on Arctic Permafrost Coastlines. Estuaries and Coasts, 35 (2), 383-400. Larson, A.M., 2011: Forest tenure reform in the age of climate change: Lessons for REDD+. Global Environmental Change, 21, 540-549. Lasco, R.D., F.B. Pulhin, P.A. Jaranilla-Sanchez, R.J.P. Delfino, R. Gerpacio, and K. Garcia, 2009: Mainstreaming adaptation in developing countries: The case of the Philippines. Climate and Development, 1 (2), 130-146. Lasco, R.D., R.V.O. Cruz, J.M. Pulhin, and F.B. Pulhin, 2010: The Case of Pantabangan-Carranglan Watershed Assessing Impacts, Vulnerability and Adaptation. Nova Science Publishers, New York, 167 pp. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Lasco, R.D., C.M.D. Habito, R.J.P. Delfino, F.B. Pulhin, and R.N. Concepcion, 2011: Climate Change Adaptation for Smallholder Farmers in Southeast Asia. World Agroforestry Centre, Laguna, Philippines, 65 pp. Lasco, R.D., R.J. Delfino, M. Rangasa, and F.B. Pulhin, 2012: The role of local government units in mainstreaming climate change adaptation: the case of Albay, Philippines. In: Local Climate Change and Society [Salih, M.A.M. (ed.)]. Routledge, London and New York, pp. 45-73. Lebel, L., Li, L., Krittasudthacheewa, Chayanis, Juntopas, M., Vijitpan, T., and Uchiyama, T., 2012: Mainstreaming climate change adaptation into development planning. Adaptation Knowledge Platform and Stockholm Environment Institute, Bangkok, 15 pp. Lee, I.-M., S.-S. Tsai, C.-K. Ho, H.-F. Chiu, and C.Y. Yang, 2007: Air pollution and hospital admissions for congestive heart failure in a tropical city: Kaohsiung, Taiwan. Inhalation Toxicology, 19 (10), 899-904. Lettenmaier, D.P., V. Aizen, A. Amani, T. Bohn, F. Giorgi, S. Harrison, T.G. Huntington, R. Lawford, P. Letitre, H. Lins, J. Magomi, G.-K. Park, I. Severskiy, W.J. Shuttleworth, P. Singh, S. Sorooshian, W. Struckmeier, K. Takeuchi, L. Tallaksen, C. Vörösmarty, T. Yan, and T. Zhang, 2009: Changes in the global water cycle, Chapter 11. In: The United Nations World Water Development Report 3: Water in a Changing World [World Water Assessment Programme (ed.)]. UNESCO and Earthscan, Paris and London, pp. 181-225. Levy, J.S. and N.C. Ban, 2013: A method for incorporating climate change modelling into marine conservation planning: An Indo-west Pacific example. Marine Policy, 38 (0), 16-24. Li, H., 2008: The more severe climate change, the greater the negative impact. China Forestry Industry, 4, 60-63. Li, J., F. Chen, E.R. Cook, X. Goua, and Y. Zhang, 2007: Drought reconstruction for north central China from tree rings: the value of the Palmer drought severity index. International Journal of Climatology, 27 (7), 903-909. Li, J.B., X.H. Gou, E.R. Cook, and F.H. Chen, 2006: Tree-ring based drought reconstruction for the central Tien Shan area in northwest China. Geophysical Research Letters, 33 (7), 5 pp. Li, R., H. Tian, and X. Li, 2010: Climate change induced range shifts of Galliformes in China. Integrative Zoology, 5 (2), 154-163. Li, X., G. Cheng, H. Jin, E. Kang, T. Che, R. Jin, L. Wu, Z. Nan, J. Wang, and Y. Shen, 2008: Cryospheric change in China. Global and Planetary Change, 62 (3 4), 210-218. Li, X., H.S. He, Z. Wu, Y. Liang, and J.E. Schneiderman, 2013: Comparing effects of climate warming, fire, and timber harvesting on a boreal forest landscape in Northeastern China. PloS One, 8 (4), 11 pp. Li, Z.-S., Q.-B. Zhang, and K. Ma, 2012: Tree-ring reconstruction of summer temperature for A.D. 1475 2003 in the central Hengduan Mountains, Northwestern Yunnan, China. Climatic Change, 110 (1), 455-467. Lian, K.K. and L. Bhullar, 2011: Governance on Adaptation to Climate Change in the Asean Region. Carbon and Climate Change Law Review, 5 (1), 82-90. Liancourt, P., L.A. Spence, B. Boldgiv, A. Lkhagva, B.R. Helliker, B.B. Casper, and P.S. Petraitis, 2012: Vulnerability of the northern Mongolian steppe to climate change: insights from flower production and phenology. Ecology, 93 (4), 815-824. Liang, T., Q. Feng, H. Yu, X. Huang, H. Lin, S. An, and J. Ren, 2012: Dynamics of natural vegetation on the Tibetan Plateau from past to future using a comprehensive and sequential classification system and remote sensing data. Grassland Science, 58 (4), 208-220. Lim, B., E. Spanger-Siegfried, I. Burton, E. Malone, and S. Huq (eds.), 2005: Adaptation Policy Frameworks for Climate Change: Developing Strategies, Policies and Measures. Cambridge University Press, New York, 258 pp. Lin, H., B. Xu, Y. Chen, and W. Wang, 2009: Legionella pollution in cooling tower water of air conditioning systems in Shanghai, China. Journal of Applied Microbiology, 106 (2), 606-612. Lioubimtseva, E. and G.M. Henebry, 2009: Climate and environmental change in arid Central Asia: Impacts, vulnerability, and adaptations. Journal of Arid Environments, 73 (11), 963-977. Liu, H., C.-L. Feng, Y.-B. Luo, B.-S. Chen, Z.-S. Wang, and H.-Y. Gu, 2010: Potential challenges of climate change to orchid conservation in a wild orchid hotspot in Southwestern China. Botanical Review, 76 (2), 174-192. Liu, H., F. Tian, H.C. Hu, H.P. Hu, and M. Sivapalan, 2013a: Soil moisture controls on patterns of grass green-up in Inner Mongolia: an index based approach. Hydrology and Earth System Sciences, 17 (2), 805-815. Liu, H., A. Park Williams, C.D. Allen, D. Guo, X. Wu, O.A. Anenkhonov, E. Liang, D.V. Sandanov, Y. Yin, Z. Qi, and N.K. Badmaeva, 2013b: Rapid warming accelerates tree growth decline in semi-arid forests of Inner Asia. Global Change Biology, 19 (8), 2500-2510. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Liu, Z., J. Yang, Y. Chang, P.J. Weisberg, and H.S. He, 2012: Spatial patterns and drivers of fire occurrence and its future trend under climate change in a boreal forest of Northeast China. Global Change Biology, 18 (6), 2041- 2056. Lloyd, A.H. and A.G. Bunn, 2007: Responses of the circumpolar boreal forest to 20th century climate variability. Environmental Research Letters, 2 (4), 14 pp. Lloyd, A.H., A.G. Bunn, and L. Berner, 2011: A latitudinal gradient in tree growth response to climate warming in the Siberian taiga. Global Change Biology, 17 (5), 1935-1945. Loucks, C., S. Barber-Meyer, M. Hossain, A. Barlow, and R. Chowdhury, 2010: Sea level rise and tigers: predicted impacts to Bangladesh s Sundarbans mangroves. Climatic Change, 98 (1), 291-298. Lucht, W., S. Schaphoff, T. Erbrecht, U. Heyder, and W. Cramer, 2006: Terrestrial vegetation redistribution and carbon balance under climate change. Carbon Balance and Management, 1, 6 pp. Ma, Y.F., L.G. Bian, C.D. Xiao, I. Allison, and X.J. Zhou, 2010: Near surface climate of the traverse route from Zhongshan Station to Dome A, East Antarctica. Antarctic Science, 22 (4), 443-459. Macchi, M., G. Oviedo, S. Gotheil, K. Cross, A. Boedhihartono, C. Wolfangel, and M. Howell, 2008: Indigenous and Traditional Peoples and Climate Change. IUCN, Gland, Switzerland, 66 pp. Mainuddin, M., M. Kirby, and C.T. Hoanh, 2011: Adaptation to climate change for food security in the lower Mekong Basin. Food Security, 3 (4), 433-450. Majra, J. and A. Gur, 2009: Climate change and health: Why should India be concerned? Indian Journal of Occupational and Environmental Medicine, 13 (1), 11. Mandych, .F., T.V. Yashina, I.A. Artemov, V.V. Dekenov, G.E. Insarov, O.V. Ostanin, I.N. Rotanova, M.G. Sukhova, N.F. Kharlamova, A.S. Shishikin, and A.B. Shmakin, 2012: Biodiversity Conservation in the Russian Portion of the Altai-Sayan Ecoregion under Climate Change. Adaptation Strategy. Gorod Publishing House, Krasnoyarsk, 62 pp. Marchenko, S.S., A.P. Gorbunov, and V.E. Romanovsky, 2007: Permafrost warming in the Tien Shan Mountains, Central Asia. Global and Planetary Change, 56 (3-4), 311-327. Markandya A, A.B., Hales S, Chiabai A, Criqui P, Mima S, Tonne C, and Wilkinson P, 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: low-carbon electricity generation. Lancet, 374 (9706), 2006- 2015. Marques, A., M.L. Nunes, S.K. Moore, and M.S. Strom, 2010: Climate change and seafood safety: human health implications. Food Research International, 43 (7), 1766-1779. Marzeion, B., A.H. Jarosch, and M. Hofer, 2012: Past and future sea-level change from the surface mass balance of glaciers. Cryosphere, 6 (6), 1295-1322. Masutomi, Y., K. Takahashi, H. Harasawa, and Y. Matsuoka, 2009: Impact assessment of climate change on rice production in Asia in comprehensive consideration of process/parameter uncertainty in general circulation models. Agriculture, Ecosystems & Environment, 131 (3 4), 281-291. Mathy, S. and C. Guivarch, 2010: Climate policies in a second-best world-A case study on India. Energy Policy, 38, 1519-1528. Mattoo, A. and A. Subramanian, 2012: Equity in climate change: An analytical review. World Development 40 (6), 1083-1097. Maxwell, J.F., 2009: Vegetation and vascular flora of the Mekong River, Kratie and Steung Treng Provinces, Cambodia. Maejo International Journal of Science and Technology, 3 (No.1), 143-211. McClanahan, T.R., S.D. Donner, J.A. Maynard, M.A. MacNeil, N.A.J. Graham, J. Maina, A.C. Baker, J.B.I. Alemu, M. Beger, S.J. Campbell, E.S. Darling, C.M. Eakin, S.F. Heron, S.D. Jupiter, C.J. Lundquist, E. McLeod, P.J. Mumby, M.J. Paddack, E.R. Selig, and R. van Woesik, 2012: Prioritizing key resilience indicators to support coral reef management in a changing climate. PloS One, 7 (8), 7 pp. McConkey, K.R., S. Prasad, R.T. Corlett, A. Campos-Arceiz, J.F. Brodie, H. Rogers, and L. Santamaria, 2012: Seed dispersal in changing landscapes. Biological Conservation, 146 (1), 1-13. McGuire, A.D., F.S. Chapin, C. Wirth, M. Apps, J. Bhatti, T. Callaghan, T.R. Christensen, J.S. Clein, M. Fukuda, T. Maximov, A. Onuchin, A. Shvidenko, and E.A. Vaganov, 2007: Responses of high latitude ecosystems to global change: Potential consequences for the climate system. In: Terrestrial Ecosystems in a Changing World [Canadell, J.G., D.E. Pataki, and L.F. Pitelka (eds.)]. Springer, Berlin, pp. 297-310. McLeod, E., K.R.N. Anthony, A. Andersson, R. Beeden, Y. Golbuu, J. Kleypas, K. Kroeker, D. Manzello, R.V. Salm, H. Schuttenberg, and J.E. Smith, 2013: Preparing to manage coral reefs for ocean acidification: lessons from coral bleaching. Frontiers in Ecology and the Environment, 11 (1), 20-27. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McMichael, A.J., P. Wilkinson, R.S. Kovats, S. Pattenden, S. Hajat, B. Armstrong, N. Vajanapoom, E.M. Niciu, H. Mahomed, and C. Kingkeow, 2008: International study of temperature, heat and urban mortality: the ISOTHURM project. International Journal of Epidemiology, 37 (5), 1121-1131. Mearns, R. and A. Norton, 2010: Social Dimensions of Climate Change: Equity and Vulnerability in a Warming World. World Bank, Washington DC, 348 pp. Meleshko, V.P. and S.M. Semenov, 2008: Assessment Report on Climate Change and Its Consequences in the Russian Federation. Moscow, 24 pp. Menon, S., M.Z. Islam, and A.T. Peterson, 2009: Projected climate change effects on nuthatch distribution and diversity across Asia. Raffles Bulletin of Zoology, 57 (2), 569-575. Merrey, D.J., P. Drechsel, F.W.P. de Vries, and H. Sally, 2005: Integrating livelihoods into integrated water resources management: taking the integration paradigm to its logical next step for developing countries. Regional Environmental Change, 5 (4), 197-204. Miettinen, J., C. Shi, and S.C. Liew, 2011a: Deforestation rates in insular Southeast Asia between 2000 and 2010. Global Change Biology, 17 (7), 2261-2270. Miettinen, J., C.H. Shi, and S.C. Liew, 2011b: Influence of peatland and land cover distribution on fire regimes in insular Southeast Asia. Regional Environmental Change, 11 (1), 191-201. MNRE, 2010: Malaysia s Second National Communication (Nc2) Submitted to the United Nations Framework Convention on Climate Change (UNFCCC). Ministry of Natural Resources and Environment (MNRE), Malaysia, 145 pp. Moench, M., 2010: Responding to climate and other change processes in complex contexts: Challenges facing development of adaptive policy frameworks in the Ganga Basin. Technological Forecasting and Social Change, 77 (6), 975-986. Mohammat, A., X. Wang, X. Xu, L. Peng, Y. Yang, X. Zhang, R.B. Myneni, and S. Piao, 2013: Drought and spring cooling induced recent decrease in vegetation growth in Inner Asia. Agricultural and Forest Meteorology, 178, 21-30. Mohammed, A.R. and L. Tarpley, 2009: High nighttime temperatures affect rice productivity through altered pollen germination and spikelet fertility. Agricultural and Forest Meteorology, 149 (6-7), 999-1008. Moiseev, P.A., A.A. Bartysh, and Z.Y. Nagimov, 2010: Climate changes and tree stand dynamics at the upper limit of their growth in the North Ural mountains. Russian Journal of Ecology, 41 (6), 486-497. Molle, F. and C.T. Hoanh, 2009: Implementing Integrated River Basin Management: Lessons from the Red River Basin, Vietnam. IWMI Research Report 131, International Water Management Institute, Colombo, Sri Lanka, 33 pp. Mondal, P., 2012: Baseline assessments, vulnerability analysis and finding sustainable livelihood options: Designing a climate change adaptation project in Ben Tre province, Vietnam. Asian Journal of Environment and Disaster Risk Management, 4 (4), 485-504. Morioka, I., N. Miyai, and K. Miyashita, 2006: Hot environment and health problems of outdoor workers at a construction site. Industrial Health, 44 (3), 474-480. Moser, S.C. and J.A. Ekstrom, 2010: A framework to diagnose barriers to climate change adaptation. Proceedings of the National Academy of Sciences of the United States of America, 107 (51), 22026-22031. MRC, 2009: Adaptation to Climate Change in the Countries of the Lower Mekong Basin: Regional Synthesis Report. MRC Technical Paper No. 24, Mekong River Commission, Vientiane, Laos, 89 pp. Mulligan, M., M. Fisher, B. Sharma, Z. Xu, C. Ringler, G. Mahé, A. Jarvis, J. Ramírez, J.-C. Clanet, and A. Ogilvie, 2011: The nature and impact of climate change in the Challenge Program on Water and Food (CPWF) basins. Water International, 36 (1), 96-124. Mumby, H.S., A. Courtiol, K.U. Mar, and V. Lummaa, 2013: Climatic variation and age-specific survival in Asian elephants from Myanmar. Ecology, 94 (5), 1131-1141. Munslow, B. and T. O'Dempsey, 2010: Globalisation and climate change in Asia: the urban health impact. Third World Quarterly, 31 (8), 1339 - 1356. Murdiyarso, D. and L. Lebel, 2007: Local to global perspectives on forest and land fires in Southeast Asia. Mitigation and Adaptation Strategies for Global Change, 12 (1), 3-11. Murphy, T.I. and M. K. Sampson, 2013: The stress of climate change on water management in Cambodia with a focus on rice productiont. Climate and Development, 5 (1), 77-92. Murthy, I.K., R. Tiwari, and N.H. Ravindranath, 2011: Climate change and forests in India: adaptation opportunities and challenges. Mitigation and Adaptation Strategies for Global Change, 16 (2), 161-175. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Murty, U.S., M.S. Rao, and N. Arunachalam, 2010: The effects of climatic factors on the distribution and abundance of Japanese encephalitis vectors in Kurnool district of Andhra Pradesh, India. Journal of Vector Borne Diseases, 47 (1), 26-32. Myers-Smith, I.H., D.S. Hik, C. Kennedy, D. Cooley, J.F. Johnstone, A.J. Kenney, and C.J. Krebs, 2011: Expansion of canopy-forming willows over the twentieth century on Herschel Island, Yukon Territory, Canada. Ambio, 40 (6), 610-623. Nadyozhina, E.D., T.V. Pavlova, I.M. Shkolnik, E.K. Molkentin, and A.A. Semioshina, 2010: Simulation of snowcover and permafrost in Russia. Earth Cryosphere, 2, 87-97. Nag, P.K., A. Nag, and S.P. Ashtekar, 2007: Thermal limits of men in moderate to heavy work in tropical farming. Industrial Health, 45 (1), 107-117. Nagai, S., G. Yoshida, and K. Tarutani, 2011: Change in species composition and distribution of algae in the coastal waters of western Japan. In: Global Warming Impacts - Case Studies on the Economy, Human Health, and on Urban and Natural Environments. [Casalegno, S. (ed.)]. InTech, Shanghai, pp. 209-237. Nakaegawa, T., A. Kitoh, and M. Hosaka, 2013: Discharge of major global rivers in the late 21st century climate projected with the high horizontal resolution MRI-AGCMs. Hydrological Processes, 1-18. Narama, C., A. Kaab, M. Duishonakunov, and K. Abdrakhmatov, 2010: Spatial variability of recent glacier area changes in the Tien Shan Mountains, Central Asia, using Corona (similar to 1970), Landsat (similar to 2000), and ALOS (similar to 2007) satellite data. Global and Planetary Change, 71 (1-2), 42-54. Neo, L., 2012: Governance issues in climate change adaptation in the Lower Mekong Basin: perspectives from practitioners. Asian Journal of Environment and Disaster Risk Management, 4 (4), 397-424. Ngoundo, M., C.E. Kan, Y.C. Chang, S.L. Tsai, and I. Tsou, 2007: Options for water saving in tropical humid and semi-arid regions using optimum compost application rates. Irrigation and Drainage, 56 (1), 87-98. Nguyen, H.N., 2007: Floods in Mekong River Delta, Viet Nam. Occasional Paper. Human Development Report 2007/2008. Fighting Climate Change: Human Solidarity in a Changing World. United Nations Development Program. Nguyen, K.D.T., S.A. Morley, C.-H. Lai, M.S. Clark, K.S. Tan, A.E. Bates, and L.S. Peck, 2011: Upper temperature limits of tropical marine ectotherms: Global warming implications. PloS One, 6 (12), 8 pp. Ni, J.A., 2011: Impacts of climate change on Chinese ecosystems: key vulnerable regions and potential thresholds. Regional Environmental Change, 11 (Suppl1), S49-S64. Nicholls, R.J., S. Hanson, C. Herweijer, N. Patmore, S. Hallegatte, J.C.-. Morlot, J. Chateau, and R. Muir-Wood, 2008: Ranking Port Cities with High Exposure and Vulnerability to Climate Extremes: Exposure Estimates. OECD Environment Working Papers, No. 1, OECD Publishing. doi:10.1787/011766488208. Niino, Y., 2011: Options on Land Management and Land Use for Coping with Climate Change in South Asia. In: Climate Change and Food Security in South Asia [Lal, R., M.V.K. Sivakumar, S.M.A. Faiz, A.H.M.M. Rahman, and K.R. Islam (eds.)]. Springer Ltd, New York, USA, pp. 277-294. Niu, D., D. Jiang, and F. Li, 2010: Higher education for sustainable development in China. International Journal of Sustainability in Higher Education, 11 (2), 153-162. Nomura, K. and O. Abe, 2010: Higher education for sustainable development in Japan: policy and progress. International Journal of Sustainability in Higher Education, 11 (2), 120-129. Noordwijk, M., 2010: Climate Change, Biodiversity, Livelihoods, and Sustainability in Southeast Asia. In: Moving Forward: Southeast Asia Perspectives on Climate Change and Biodiversity [Sajise, P.E., M.V. Ticsay, and G.C. Saguigut (eds.)], p. 25. Noroozi, J., H. Pauli, G. Grabherr, and S.-W. Breckle, 2011: The subnival-nival vascular plant species of Iran: a unique high-mountain flora and its threat from climate warming. Biodiversity and Conservation, 20 (6), 1319- 1338. Nuorteva, P., M. Keskinen, and O. Varis, 2010: Water, livelihoods and climate change adaptation in the Tonle Sap Lake area, Cambodia: learning from the past to understand the future. Journal of Water and Climate Change, 1, 87-101. Ogawa-Onishi, Y. and P.M. Berry, 2013: Ecological impacts of climate change in Japan: The importance of integrating local and international publications. Biological Conservation, 157, 361-371. Ohta, S. and A. Kimura, 2007: Impacts of climate changes on the temperature of paddy waters and suitable land for rice cultivation in Japan. Agricultural and Forest Meteorology, 147 (3 4), 186-198. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Okunishi, T., S.-i. Ito, T. Hashioka, T.T. Sakamoto, N. Yoshie, H. Sumata, Y. Yara, N. Okada, and Y. Yamanaka, 2012: Impacts of climate change on growth, migration and recruitment success of Japanese sardine (Sardinops melanostictus) in the western North Pacific. Climatic Change, 115 (3-4), 485-503. Olden, J.D., M.J. Kennard, J.J. Lawler, and N.L. Poff, 2010: Challenges and opportunities in implementing managed relocation for conservation of freshwater species. Conservation Biology, 25 (1), 40-47. Onozuka, D., M. Hashizume, and A. Hagihara, 2010: Effects of weather variability on infectious gastroenteritis. Epidemiology and Infection, 138 (2), 236. Orr, S., Pittock, J., Chapagain, A., & Dumaresq, D., 2012: Dams on the Mekong River: Lost fish protein and the implications for land and water resources. Global Environmental Change, 22 (4), 925-932. Ortiz, R., K.D. Sayre, B. Govaerts, R. Gupta, G.V. Subbarao, T. Ban, D. Hodson, J.M. Dixon, J. Iván Ortiz- Monasterio, and M. Reynolds, 2008: Climate change: Can wheat beat the heat? Agriculture, Ecosystems & Environment, 126 (1 2), 46-58. Osawa, A., Y. Matsuura, and T. Kajimoto, 2010: Characteristics of permafrost forests in Siberia and potential responses to warming climate. In: Permafrost Ecosystems: Siberian Larch Forests [Osawa, A., O.A. Zyryanova, Y. Matsuura, T. Kajimoto, and R.W. Wein (eds.)]. Springer, Berlin, pp. 459-481. Paaijmans, K.P., S. Blanford, B.H. Chan, and M.B. Thomas, 2012: Warmer temperatures reduce the vectorial capacity of malaria mosquitoes. Biology Letters, 8 (3), 465-468. Page, S.E., J.O. Rieley, and C.J. Banks, 2011: Global and regional importance of the tropical peatland carbon pool. Global Change Biology, 17 (2), 798-818. Pal, I. and A. Al-Tabbaa, 2009: Trends in seasonal precipitation extremes - An indicator of 'climate change' in Kerala, India. Journal of Hydrology, 367 (1-2), 62-69. Panday, P.K. and B. Ghimire, 2012: Time-series analysis of NDVI from AVHRR data over the Hindu Kush- Himalayan region for the period 1982-2006. International Journal of Remote Sensing, 33 (21), 6710-6721. Panyakul, V.R., 2012: Climate change adaptation through agro-social enterprise: Green nets experiences in Thailand. Asian Journal of Environment and Disaster Management, 4 (4), 511-526. Park, J.H., L. Duan, B. Kim, M.J. Mitchell, and H. Shibata, 2010: Potential effects of climate change and variability on watershed biogeochemical processes and water quality in Northeast Asia. Environment International, 36 (2), 212-225. Partridge, J., P. Ghimire, T. Sedai, M.B. Bista, and M. Banerjee, 2007: Endemic Japanese encephalitis in the Kathmandu valley, Nepal. American Journal of Tropical Medicine and Hygiene, 77 (6), 1146-1149. Pascual, M., X. Rodó, S. P. Ellner, R. Colwell, and M. J. Bouma, 2000: Cholera dynamics and El Nino-Southern Oscillation. Science, 289, 1766-1769. Patnaik, U. and K. Narayanan, 2009: Vulnerability and climate change: An analysis of the eastern coastal districts of India. MPRA Paper No. 22062, Munich Personal RePEc Archive, 19 pp. Paul, H., A. Ernsting, S. Semino, S. Gura, and A. Lorch, 2009: Agriculture and climate change: Real problems, false solutions. Econexus, Biofuelwatch, Grupo de Reflexion Rural and NOAH - Friends of the Earth, Bangkok. Paul, S.K. and J.K. Routray, 2010: Flood proneness and coping strategies: the experiences of two villages in Bangladesh. Disasters, 34 (2), 489-508. Pavlidis, Y.A., S.L. Nikiforov, S.A. Ogorodov, and G.A. Tarasov, 2007: The Pechora sea: Past, recent, and future. Oceanology, 47 (6), 865-876. Pawar, A., R. Bansal, M. Kumar, N. Jain, and K. Vaishnav, 2008: A rapid assessment of mosquito breeding, vector control measures and treatment seeking behaviour in selected slums of Surat, Gujarat, India, during post-flood period. Journal of Vector Borne Diseases, 45 (4), 325-327. Paz, S., N. Bisharat, E. Paz, O. Kidar, and D. Cohen, 2007: Climate change and the emergence of Vibrio vulnificus disease in Israel. Environmental Research, 103 (3), 390-396. Pearson, R.G., S.J. Phillips, M.M. Loranty, P.S.A. Beck, T. Damoulas, S.J. Knight, and S.J. Goetz, 2013: Shifts in Arctic vegetation and associated feedbacks under climate change. Nature Climate Change, 3 (7), 673-677. Peh, K.S.H., M.C.K. Soh, N.S. Sodhi, W.F. Laurance, D.J. Ong, and R. Clements, 2011: Up in the clouds: Is sustainable use of tropical montane cloud forests possible in Malaysia? Bioscience, 61 (1), 27-38. Peng, S., A. Chen, L. Xu, C. Cao, J. Fang, R.B. Myneni, J.E. Pinzon, C.J. Tucker, and S. Piao, 2011: Recent change of vegetation growth trend in China. Environmental Research Letters, 6 (4), 044027. Penning-Rowsell, E., P. Sultana, and P. Thompson, 2011: Migration and Global Environmental Change: Population Movement in Response to Climate-Related Hazards in Bangladesh: The Last Resort . Flood Hazard Research Centre, Middlesex University, London, UK, 38 pp. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Peras, R.J.J., J.M. Pulhin, R.D. Lasco, R.V.O. Cruz, and F.B. Pulhin, 2008: Climate variability and extremes in the Pantabangan-Carranglan Watershed, Philippines: Assessment of impacts and adaptation practices. Journal of Environmental Science and Management, 11 (2), 14-31. Persha, L., H. Fischer, A. Chhatre, A. Agrawal, and C. Benson, 2010: Biodiversity conversion and livelihoods in human-dominated landscapes: Forest commons in South Asia. Biological Conservation, 143, 2918-2925. Pfister, S., A. Koehler, and S. Hellweg, 2009: Assessing the environmental impacts of freshwater consumption in LCA. Environmental Science & Technology, 43 (11), 4098-4104. Post, E., U.S. Bhatt, C.M. Bitz, J.F. Brodie, T.L. Fulton, M. Hebblewhite, J. Kerby, S.J. Kutz, I. Stirling, and D.A. Walker, 2013: Ecological consequences of sea-ice decline. Science, 341 (6145), 519-524. Poulter, B., N. Pederson, H. Liu, Z. Zhu, R. D'Arrigo, P. Ciais, N. Davi, D. Frank, C. Leland, R. Myneni, S. Piao, and T. Wang, 2013: Recent trends in inner asian forest dynamics to temperature and precipitation indicate high sensitivity to climate change. Agricultural and Forest Meteorology, 178 (15), 31-45. Prabhakar, S.V.R.K., T. Kobashi, and A. Srinivasan, 2010: Monitoring Progress of Adaptation to Climate Change: The Use of Adaptation Metrics. Asian Journal of Environment and Disaster Management 2(3), 435-442. Prabhakar, S.V.R.K. and A. Srinivasan, 2011: Metrics for Mainstreaming Adaptation in Agriculture Sector. In: [Lal, R., M.V.K. Sivakumar, S.M.A. Faiz, A.H.M.M. Rahman, and K.R. Islam (eds.)]. Springer, New York, USA, pp. 551-568. PRB, 2012: World Population Data Sheet 2012. Population Reference Bureau, Washington DC, USA, 20 pp. Prentice, M.L. and S. Glidden, 2010: Glacier crippling and the rise of the snowline in western New Guinea (Papua Province, Indonesia) from 1972 to 2000. In: Altered Ecologies: Fire, Climate and Human Influence on Terrestrial Landscapes [Haberle, S.G., J. Stevenson, and M. Prebble (eds.)], pp. 457-471. Putz, F.E., P.A. Zuidema, T. Synnott, M. Pena-Claros, M.A. Pinard, D. Sheil, J.K. Vanclay, P. Sist, S. Gourlet- Fleury, B. Griscom, J. Palmer, and R. Zagt, 2012: Sustaining conservation values in selectively logged tropical forests: The attained and the attainable. Conservation Letters, 5 (4), 296-303. Qian, Y., S. Li, Q. Wang, K. Yang, G. Yang, S. Lü, and X. Zhou, 2010: Advances on impact of climate change on human health. Advances in Climate Change Research, 6 (4), 241-247. Qin, Z., Q. Zhuang, X. Zhu, X. Cai, and X. Zhang, 2011: Carbon consequences and agricultural implications of growing biofuel crops on marginal agricultural lands in China. Environmental Science & Technology, 45 (24), 10765-10772. Qiu, Y.S., Z.J. Lin, and Y.Z. Wang, 2010: Responses of fish production to fishing and climate variability in the northern South China Sea. Progress in Oceanography, 85 (3-4), 197-212. Räsänen, T.A.K., J. Lauri, H. and Kummu, M. , 2012: Downstream hydrological impacts of hydropower development in the Upper Mekong Basin. Water Resources Management, 26 (12), 3495-3513. Radiæ, V., A. Bliss, A.C. Beedlow, R. Hock, E. Miles, and J.G. Cogley, 2013: Regional and global projections of twenty-first century glacier mass changes in response to climate scenarios from global climate models. Climate Dynamics, 1-22. Ranger, N., S. Hallegatte, S. Bhattacharya, M. Bachu, S. Priya, K. Dhore, F. Rafique, P. Mathur, N. Naville, F. Henriet, C. Herweijer, S. Pohit, and J. Corfee-Morlot, 2011: An assessment of the potential impact of climate change on flood risk in Mumbai. Climatic Change, 104 (1), 139-167. Ratnakumar, P., V. Vadez, L. Krishnamurthy, and G. Rajendrudu, 2011: Semi-arid crop responses to atmospheric elevated CO2. Plant Stress, 5 (1), 42-51. Ravallion, M., S. Chen, and P. Sangraula, 2007: New Evidence on the Urbanization of Global Poverty. World Bank, Washington D.C. Razumov, S.O., 2010: Permafrost as a factor of the dynamics of the coastal zone of the Russian East Arctic Seas. Oceanology, 50 (2), 262-267. Ren, W., H. Tian, B. Tao, A. Chappelka, G. Sun, C. Lu, M. Liu, G. Chen, and X. Xu, 2011: Impacts of tropospheric ozone and climate change on net primary productivity and net carbon exchange of China's forest ecosystems. Global Ecology and Biogeography, 20 (3), 391-406. Revi, A., 2008: Climate change risk: an adaptation and mitigation agenda for Indian cities. Environment and Urbanization, 20 (1), 207-229. Reynolds, C.C.O. and M. Kandikar, 2008: Climate impacts of air quality policy: Switching to a natural gas-fueled public transportation system in New Delhi. Environmental Science & Technology, 42 (16), 5860-5865. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Richardson, A.D., T.F. Keenan, M. Migliavacca, Y. Ryu, O. Sonnentag, and M. Toomey, 2013: Climate change, phenology, and phenological control of vegetation feedbacks to the climate system. Agricultural and Forest Meteorology, 169 (0), 156-173. Riseborough, D., N. Shiklomanov, B. Etzelmüller, S. Gruber, and S. Marchenko, 2008: Recent advances in permafrost modelling. Permafrost and Periglacial Processes, 19 (2), 137-156. Rodó, X., M. Pascual, G. Fuchs, and A. Faruque, 2002: ENSO and cholera: A nonstationary link related to climate change? Proceedings of the National Academy of Sciences of the United States of America, 99 (20), 12901- 12906. Rodell, M., I. Velicogna, and J.S. Famiglietti, 2009: Satellite-based estimates of groundwater depletion in India. Nature, 460 (7258), 999-1002. Romanovsky, V.E., A.L. Kholodov, S.S. Marchenko, N.G. Oberman, D.S. Drozdov, G.V. Malkova, N.G. Moskalenko, A.A. Vasiliev, D.O. Sergeev, and M.N. Zheleznyak, 2008: Thermal State and Fate of Permafrost in Russia: First results of IPY. In: Ninth International Conference on Permafrost, Vol. 1 [Kane, D.L., and K.M. Hinkel (eds.)]. Proceedings of the Ninth International Conference on Permafrost, June 29 - July 3, 2008, pp.1511-1518. Romanovsky, V.E., D.S. Drozdov, N.G. Oberman, G.V. Malkova, A.L. Kholodov, S.S. Marchenko, N.G. Moskalenko, D.O. Sergeev, N.G. Ukraintseva, A.A. Abramov, D.A. Gilichinsky, and A.A. Vasiliev, 2010: Thermal state of permafrost in Russia. Permafrost and Periglacial Processes, 21 (2), 136-155. Romero-Lankao P, W.T., and Gnatz D. , 2011: Conclusions and Policy Directions. In: Cities and Climate Change, Global Report on Human Settlements 2011. Earthscan, London, Washington DC, pp. 163-183. Rosegrant, M.W., 2011: Impacts of Climate Change on Food Security and Livelihoods. Food Security and Climate Change in Dry Areas: Proceedings of an International Conference, 1-4 February 2010, Amman, Jordan,24-26. Roth, C.H. and C.M. Grunbuhel, 2012: Developing multi-scale adaptation strategies: a case study for farming communities in Cambodia and Laos. Asian Journal of Environment and Disaster Risk Management, 4 (4), 441- 462. Ryan, A., D. Tilbury, P.B. Corcoran, O. Abe, and K. Nomura, 2010: Sustainability in higher education in the Asia- Pacific: developments, challenges and prospects. International Journal of Sustainability in Higher Education, 11 (2), 106-119. Sadoff, C. and M. Muller, 2009: Water Management, Water Security and Climate Change Adaptation: Early Impacts and Essential Responses. Technical Committee (TEC) Background Paper, No.14, Global Water Partnership, Stockholm, Sweden, 92 pp. Sakai, S., R.D. Harrison, K. Momose, K. Kuraji, H. Nagamasu, T. Yasunari, L. Chong, and T. Nakashizuka, 2006: Irregular droughts trigger mass flowering in aseasonal tropical forests in Asia. American Journal of Botany, 93 (8), 1134-1139. Salick, J., Z. Fang, and A. Byg, 2009: Eastern Himalayan alpine plant ecology, Tibetan ethnobotany, and climate change. Global Environmental Change, 19 (2), 147-155. Salick, J. and N. Ross, 2009: Traditional peoples and climate change. Global Environmental Change, 19 (2), 137- 139. Sano, M., F. Furuta, and T. Sweda, 2010: Summer temperature variations in southern Kamchatka as reconstructed from a 247-year tree-ring chronology of Betula ermanii. Journal of Forest Research, 15 (4), 234-240. Sasaki, N., G.P. Asner, W. Knorr, P.B. Durst, H.R. Priyadi, and F.E. Putz, 2011: Approaches to classifying and restoring degraded tropical forests for the anticipated REDD+ climate change mitigation mechanism. iForest- Biogeosciences and Forestry, 4 (1), 1-6. Sato, T., F. Kimura, and A. Kitoh, 2007: Projection of global warming onto regional precipitation over Mongolia using a regional climate model. Journal of Hydrology, 333 (1), 144-154. Sato, Y., T. Kojiri, Y. Michihiro, Y. Suzuki, and E. Nakakita, 2012: Estimates of climate change impact on river discharge in Japan based on a super-high-resolution climate model. Terrestrial Atmospheric and Oceanic Sciences, 23 (5), 527-540. Sato, Y., T. Kojiri, Y. Michihiro, Y. Suzuki, and E. Nakakita, 2013: Assessment of climate change impacts on river discharge in Japan using the super-high-resolution MRI-AGCM. Hydrological Processes, 1-16. Schaefer, K., T.J. Zhang, L. Bruhwiler, and A.P. Barrett, 2011: Amount and timing of permafrost carbon release in response to climate warming. Tellus Series B-Chemical and Physical Meteorology, 63 (2), 165-180. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Schaffer, A.S. and L. Ding, 2012: Strengthening climate adaptation in the Lower Mekong River basin through a regional adaptation action network. Asian Journal of Environment and Disaster Risk Management, 4 (4), 543- 565. Schlüter, M., D. Hirsch, and C. Pahl-Wostl, 2010: Coping with change: responses of the Uzbek water management regime to socio-economic transition and global change. Environmental Science & Policy, 13 (7), 620-636. Selvaraju, R., A.R. Subbiah, S. Baas, and I. Juergens, 2006: Livelihood Adaptation to Climate Variability and Change in Drought-prone Areas of Bangladesh: Developing Institutions and Options. Food and Agriculture Organization, Rome. Shahgedanova, M., G. Nosenko, T. Khromova, and A. Muraveyev, 2010: Glacier shrinkage and climatic change in the Russian Altai from the mid-20th century: An assessment using remote sensing and PRECIS regional climate model. Journal of Geophysical Research-Atmospheres, 115, D16107. Shahgedanova, M., G. Nosenko, I. Bushueva, and M. Ivanov, 2012: Changes in area and geodetic mass balance of small glaciers, Polar Urals, Russia, 1950-2008. Journal of Glaciology, 58 (211), 953-964. Shang, C.-S., C.-T. Fang, C.-M. Liu, T.-H. Wen, K.-H. Tsai, and C.-C. King, 2010: The role of imported cases and favorable meteorological conditions in the onset of dengue epidemics. PLoS Neglected Tropical Diseases, 4 (8), e775. Shao, X., Y. Xu, Z.Y. Yin, E. Liang, H. Zhu, and S. Wang, 2010: Climatic implications of a 3585-year tree-ring width chronology from the northeastern Qinghai-Tibetan Plateau. Quaternary Science Reviews, 29 (17-18), 2111-2122. Sharkhuu, N., A. Sharkhuu, V.E. Romanovsky, K. Yoshikawa, F.E. Nelson, and N.I. Shiklomanov, 2008: Thermal State of Permafrost in Mongolia. In: Ninth International Conference on Permafrost, Vol. 1 [Kane, D.L., and K.M. Hinkel (eds.)]. Proceedings of the Ninth International Conference on Permafrost, June 29 - July 3, 2008, pp.1633-1638. Shaw, R., F. Mallick, and Y. Takeuchi, 2011: Essentials of Higher Education in Disaster Risk Reduction: Prospects and Challenges. Disaster Education. Emerald Publisher, UK. Shen, M., Z. Sun, S. Wang, G. Zhang, W. Kong, A. Chen, and S. Piao, 2013: No evidence of continuously advanced green-up dates in the Tibetan Plateau over the last decade. Proceedings of the National Academy of Sciences of the United States of America, 110 (26), E2329-E2329. Shishov, V.V. and E.A. Vaganov, 2010: Dendroclimatological evidence of climate changes across Siberia. In: Environmental Change in Siberia: Earth Observation, Field Studies and Modelling [Balzter, H. (ed.)]. Springer, Netherlands, pp. 101-114. Shoo, L.P., C. Storlie, J. Vanderwal, J. Little, and S.E. Williams, 2011: Targeted protection and restoration to conserve tropical biodiversity in a warming world. Global Change Biology, 17 (1), 186-193. Shrestha, A.B. and R. Aryal, 2011: Climate change in Nepal and its impact on Himalayan glaciers. Regional Environmental Change, 11, S65-S77. Shrestha, U.B., S. Gautam, and K.S. Bawa, 2012: Widespread climate change in the Himalayas and associated changes in local ecosystems. PloS One, 7 (5), e36741-e36741. Shuang-He, S., S.-B. Yang, Y.-X. Zhao, Y.-L. Xu, X.-Y. Zhao, Z.-Y. Wang, J. Liu, and W.-W. Zhang, 2011: Simulating the rice yield change in the middle and lower reaches of the Yangtze River under SRES B2 scenario. Acta Ecologica Sinica, 31 (1), 40-48. Siegfried, T., T. Bernauer, R. Guiennet, S. Sellars, A.W. Robertson, J. Mankin, and P. Bauer-Gottwein, 2010: Coping With International Water Conflict in Central Asia: Implications of Climate Change and Melting Ice in the Syr Darya Catchment. International Peace Research Institute, Oslo, Norway, 36 pp. Singh, C.P., S. Panigrahy, A. Thapliyal, M.M. Kimothi, P. Soni, and J.S. Parihar, 2012: Monitoring the alpine treeline shift in parts of the Indian Himalayas using remote sensing. Current Science, 102 (4), 559-562. Sitch, S., C. Huntingford, N. Gedney, P.E. Levy, M. Lomas, S.L. Piao, R. Betts, P. Ciais, P. Cox, P. Friedlingstein, C.D. Jones, I.C. Prentice, and F.I. Woodward, 2008: Evaluation of the terrestrial carbon cycle, future plant geography and climate-carbon cycle feedbacks using five Dynamic Global Vegetation Models (DGVMs). Global Change Biology, 14 (9), 2015-2039. Sivakumar, M.V. and R. Stefanski, 2011: Climate Change in South Asia. In: Climate Change and Food Security in South Asia, pp. 13-30. Skoufias, E., M. Rabassa, and S. Olivieri, 2011: The Poverty Impacts of Climate Change: A Review of the Evidence. World Bank Policy Research Working Paper 5622, World Bank, Washington DC, USA, 35 pp. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Smith, L.C. and S.R. Stephenson, 2013: New Trans-Arctic shipping routes navigable by midcentury. Proceedings of the National Academy of Sciences of the United States of America, 110 (13), E1191-E1195. Sodhi, N.S., M.R.C. Posa, T.M. Lee, D. Bickford, L.P. Koh, and B.W. Brook, 2010: The state and conservation of Southeast Asian biodiversity. Biodiversity and Conservation, 19 (2), 317-328. Sohan, L., B. Shyamal, T.S. Kumar, M. Malini, K. Ravi, V. Venkatesh, M. Veena, and S. Lal, 2008: Studies on leptospirosis outbreaks in Peddamandem Mandal of Chittoor district, Andhra Pradesh. Journal of Communicable Diseases, 40 (2), 127. Soja, A.J., N.M. Tchebakova, N.H.F. French, M.D. Flannigan, H.H. Shugart, B.J. Stocks, A.I. Sukhinin, E.I. Parfenova, F.S. Chapin, and P.W. Stackhouse, 2007: Climate-induced boreal forest change: Predictions versus current observations. Global and Planetary Change, 56 (3-4), 274-296. Solberg, K., 2010: Worst floods in living memory leave Pakistan in paralysis. The Lancet, 376 (9746), 1039-1040. Sorg, A., T. Bolch, M. Stoffel, O. Solomina, and M. Beniston, 2012: Climate change impacts on glaciers and runoff in Tien Shan (Central Asia). Nature Climate Change, 2 (10), 725-731. Spalding, M., C. Ravilious, and E.P. Green, 2001: World Atlas of Coral Reefs. University of California Press, Berkeley, 424 pp. Spotila, J.R., 2004: Sea Turtles: A Complete Guide to Their Biology, Behavior, and Conservation. Johns Hopkins University Press, Baltimore, 227 pp. Sriprom, M., K. Chalvet-Monfray, T. Chaimane, K. Vongsawat, and D. Bicout, 2010: Monthly district level risk of dengue occurrences in Sakon Nakhon Province, Thailand. Science of the Total Environment, 408 (22), 5521- 5528. Srivastava, A., S. Naresh Kumar, and P.K. Aggarwal, 2010: Assessment on vulnerability of sorghum to climate change in India. Agriculture, Ecosystems and Environment, 138 (3 4), 160-169. Stage, J., 2010: Economic valuation of climate change adaptation in developing countries. Annals of the New York Academy of Sciences, 1185, 150-163. Stewart, M.G., X.M. Wang, and M.N. Nguyen, 2012: Climate change adaptation for corrosion control of concrete infrastructure. Structural Safety, 35, 29-39. Stokes, C.R., M. Shahgedanova, I.S. Evans, and V.V. Popovnin, 2013: Accelerated loss of alpine glaciers in the Kodar Mountains, south-eastern Siberia. Global and Planetary Change, 101, 82-96. Storch, H. and N.K. Downes, 2011: A scenario-based approach to assess Ho Chi Minh City's urban development strategies against the impact of climate change. Cities, 28 (6), 517-526. Stucki, V. and M. Smith, 2011: Integrated approaches to natural resources management in practice: the catalyzing role of National Adaptation Programmes for Action. AMBIO: A Journal of the Human Environment, 40 (4), 351-360. Su, G.L.S., 2008: Correlation of climatic factors and dengue incidence in Metro Manila, Philippines. AMBIO: A Journal of the Human Environment, 37 (4), 292-294. Su, Y., Y. Weng, and Y. Chiu, 2009: Climate change and food security in East Asia. Asia Pacific Journal of Clinical Nutrition, 18 (4), 674-678. Sugiura, T., H. Sumida, S. Yokoyama, and H.Ono, 2012: Overview of recent effects of global warming on agricultural production in Japan. JARQ (Japan Agricultural Research Quarterly), 46 (1), 7-13. Sumaila, U.R., W.W.L. Cheung, V.W.Y. Lam, D. Pauly, and S. Herrick, 2011: Climate change impacts on the biophysics and economics of world fisheries. Nature Climate Change, 1 (9), 449-456. Sun, P., Z. Yu, S. Liu, X. Wei, J. Wang, N. Zegre, and N. Liu, 2012: Climate change, growing season water deficit and vegetation activity along the north-south transect of eastern China from 1982 through 2006. Hydrology and Earth System Sciences, 16 (10), 3835-3850. Surazakov, A.B., V.B. Aizen, E.M. Aizen, and S.A. Nikitin, 2007: Glacier changes in the Siberian Altai Mountains, Ob river basin, (1952 2006) estimated with high resolution imagery. Environmental Research Letters, 2. Sutton, W.R., J.P. Srivastava, and J.E. Neumann, 2013: Looking Beyond the Horizon: How Climate Change Impacts and Adaptation Responses Will Reshape Agriculture in Eastern Europe and Central Asia. World Bank Publications. Syvitski, J.P.M., A.J. Kettner, I. Overeem, E.W.H. Hutton, M.T. Hannon, G.R. Brakenridge, J. Day, C. Vörösmarty, Y. Saito, L. Giosan, and R.J. Nicholls, 2009: Sinking deltas due to human activities. Nature Geoscience, 2 (10), 681-686. Tan, J., Y. Zheng, G. Song, L.S. Kalkstein, A.J. Kalkstein, and X. Tang, 2007: Heat wave impacts on mortality in Shanghai, 1998 and 2003. International Journal of Biometeorology, 51 (3), 193-200. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Tan, J., Y. Zheng, X. Tang, C. Guo, L. Li, G. Song, X. Zhen, D. Yuan, A.J. Kalkstein, and F. Li, 2010: The urban heat island and its impact on heat waves and human health in Shanghai. International Journal of Biometeorology, 54 (1), 75-84. Tan, Z.-H., M. Cao, G.-R. Yu, J.-W. Tang, X.-B. Deng, Q.-H. Song, Y. Tang, Z. Zheng, W.-J. Liu, Z.-L. Feng, Y. Deng, J.-L. Zhang, N. Liang, and Y.-P. Zhang, 2013: High sensitivity of a tropical rainforest to water variability: Evidence from 10 years of inventory and eddy flux data. Journal of Geophysical Research: Atmospheres, 118 (16), 9393-9400. Tanaka, K., S. Taino, H. Haraguchi, G. Prendergast, and M. Hiraoka, 2012b: Warming off southwestern Japan linked to distributional shifts of subtidal canopy-forming seaweeds. Ecology and Evolution, 2 (11), 2854-2865. Tanaka, N., K. Nakao, I. Tsuyama, M. Higa, E. Nakazono, and T. Matsui, 2012a: Predicting the impact of climate change on potential habitats of fir (Abies) species in Japan and on the East Asian continent. Procedia Environmental Sciences, 13 (0), 455-466. Tao, F., Z. Zhang, J. Liu, and M. Yokozawa, 2009: Modelling the impacts of weather and climate variability on crop productivity over a large area: A new super-ensemble-based probabilistic projection. Agricultural and Forest Meteorology, 149 (8), 1266-1278. Tao, F. and Z. Zhang, 2010: Impacts of climate change as a function of global mean temperature: maize productivity and water use in China. Climatic Change, 105 (3-4), 409-432. Tao, F. and Z. Zhang, 2013a: Climate change, high-temperature stress, rice productivity, and water use in Eastern China: a new superensemble-based probabilistic projection. Journal of Applied Meteorology and Climatology, 52 (3), 531-551. Tao, F. and Z. Zhang, 2013b: Climate change, wheat productivity and water use in the North China Plain: A new super-ensemble-based probabilistic projection. Agricultural and Forest Meteorology, 170 (15), 146-165. Taylor, J., 2011: Community-based vulnerability assessment: Semarang, Indonesia. In: Resilient Cities: Cities and Adaptation to Climate Change - Proceedings of the Global Forum 2010, Local Sustainability [Otto- Zimmermann, K. (ed.)]. Springer, Dordrecht, pp. 329-337. Tchebakova, N.M., G.E. Rehfeldt, and E.I. Parfenova, 2010: From vegetation zones to climatypes: effects of climate warming on Siberian ecosystems. In: Permafrost Ecosystems: Siberian Larch Forests [Osawa, A., O.A. Zyryanova, Y. Matsuura, T. Kajimoto, and R.W. Wein (eds.)]. Springer, Berlin, pp. 427-446. Tchebakova, N.M., E.I. Parfenova, and A.J. Soja, 2011: Climate change and climate-induced hot spots in forest shifts in central Siberia from observed data. Regional Environmental Change, 11 (4), 817-827. Telles, S., N. Singh, and M. Joshi, 2009: Risk of posttraumatic stress disorder and depression in survivors of the floods in Bihar, India. Indian Journal of Medical Sciences, 63 (8), 330. Telwala, Y., B.W. Brook, K. Manish, and M.K. Pandit, 2013: Climate-induced elevational range shifts and increase in plant species richness in a Himalayan biodiversity epicentre. PloS One, 8 (2), 8 pp. Terazono, Y., Y. Nakamura, Z. Imoto, and M. Hiraoka, 2012: Fish response to expanding tropical Sargassum beds on the temperate coasts of Japan. Marine Ecology Progress Series, 464, 209-220. Thomas, R.J., 2008: Opportunities to reduce the vulnerability of dryland farmers in Central and West Asia and North Africa to climate change. Agriculture, Ecosystems & Environment, 126 (1-2), 36-45. Thomson, A.M., R.C. Izaurralde, N.J. Rosenberg, and X. He, 2006: Climate change impacts on agriculture and soil carbon sequestration potential in the Huang-Hai Plain of China. Agriculture, Ecosystems & Environment, 114 (2 4), 195-209. Thomson, A.M., K.V. Calvin, L.P. Chini, G. Hurtt, J.A. Edmonds, B. Bond-Lamberty, S. Frolking, M.A. Wise, and A.C. Janetos, 2010: Climate mitigation and the future of tropical landscapes. Proceedings of the National Academy of Sciences of the United States of America, 107, 19633-19638. Tian, X., T. Matsui, S. Li, M. Yoshimoto, K. Kobayasi, and T. Hasegawa, 2010: Heat-induced floret sterility of hybrid rice (Oryza sativa L.) cultivars under humid and low wind conditions in the field of Jianghan Basin, China. Plant Production Science, 13 (3), 243-251. Tian, Y., H. Kidokoro, T. Watanabe, Y. Igeta, H. Sakaji, and S. Ino, 2012: Response of yellowtail, Seriola quinqueradiata, a key large predatory fish in the Japan Sea, to sea water temperature over the last century and potential effects of global warming. Journal of Marine Systems, 91 (1), 1-10. Tirado, M., R. Clarke, L. Jaykus, A. McQuatters-Gollop, and J. Frank, 2010: Climate change and food safety: A review. Food Research International, 43 (7), 1745-1765. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Tischbein, B., A.M. Manschadi, A.K. Hornidge, C. Conrad, J.P.A. Lamers, L. Oberkircher, G. Schorcht, and P.L.G. Vlek, 2011: Proposals for the more efficient utilization of water resources in the Province of Khorezm, Uzbekistan. Hydrologie und Wasserbewirtschaftung, 55 (2), 116-125. Tobey, J., P. Rubinoff, D. Robadue, Jr., G. Ricci, R. Volk, J. Furlow, and G. Anderson, 2010: Practicing coastal adaptation to climate change: Lessons from integrated coastal management. Coastal Management, 38 (3), 317- 335. Tosca, M.G., J.T. Randerson, and C.S. Zender, 2012: Global impact of contemporary smoke aerosols from landscape fires on climate and the Hadley circulation. Atmospheric Chemistry and Physics Discussions, 12, 28069-28108. Tougou, D., D.L. Musolin, and K. Fujisaki, 2009: Some like it hot! Rapid climate change promotes changes in distribution ranges of Nezara viridula and Nezara antennata in Japan. Entomologia Experimentalis et Applicata, 130 (3), 249-258. Tseng, C.-T., C.-L. Sun, S.-Z. Yeh, S.-C. Chen, W.-C. Su, and D.-C. Liu, 2011: Influence of climate-driven sea surface temperature increase on potential habitats of the Pacific saury (Cololabis saira). ICES Journal of Marine Science, 68 (6), 1105-1113. Tuanmu, M.-N., A. Vina, J.A. Winkler, Y. Li, W. Xu, Z. Ouyang, and J. Liu, 2012: Climate-change impacts on understorey bamboo species and giant pandas in China's Qinling Mountains. Nature Climate Change, 3, 249- 253. Tyler, S. and M. Moench, 2012: A framework for urban climate resilience. Climate and Development, 4 (4), 311- 326. Udomratn, P., 2008: Mental health and the psychosocial consequences of natural disasters in Asia. International Review of Psychiatry, 20 (5), 441-444. UN-Habitat, 2010: The State of Asian Cities 2010/11. United Nations Human Settlements Programme, United Nations Economic and Social Commission for Asia and the Pacific, Fukuoka, Japan, 280 pp. UN-Habitat, 2011: Cities and Climate Change: Global Report on Human Settlements 2011. Earthscan, London, Washington DC, 300 pp. UN, 2009: The Millennium Development Goals Report 2009. United Nations, New York, USA, 60 pp. UN, 2012: World Urbanization Prospects: The 2011 Revision. United Nations, Department of Economic and Social Affairs, Population Division, New York, USA, 33 pp. UN, 2013: World Population Prospects: The 2012 Revision, Volume I: Comprehensive Tables ST/ESA/SER.A/336., Department of Economic and Social Affairs, Population Division, New York, USA, 463 pp. UN ESCAP, 2011: Statistical Yearbook for Asia and the Pacific 2011. United Nations, Economic and Social Commission for Asia and the Pacific, Bangkok, Thailand, 287 pp. UNDP, 2006: Human Development Report 2006. Beyond Scarcity: Power, Poverty and the Global Water Crisis. United Nations Development Programme (UNDP), New York, USA, 422 pp. UNEP, 2012: Summary for Policy Makers Highlights the Findings of the Fifth Global Environment Outlook (GEO- 5) Report. United Nations Environment Programme, Valletta, Malta, 20 pp. UNESCO, 2012: The United Nations World Water Development Report 4: Managing Water under Uncertainty and Risk. United Nations World Water Assessment Programme, United Nations Educational, Scientific and Cultural Organization, Paris, France, 397 pp. UNFCCC, 2009: Potential Costs and Benefits of Adaptation Options: A Review of Existing Literature: Technical Paper. FCCC/TP/2009/2/Rev.1, United Nation Framwork Convention on Climate Change, Bonn, Germany, 80 pp. UNISDR, 2011: Revealing Risk, Redefining Development. Global Assessment Report on Disaster Risk Reduction 2011, United Nations International Strategy for Disaster Reduction, Geneva, Switzerland, 178 pp. Uprety, K. and S.M.A. Salman, 2011: Legal aspects of sharing and management of transboundary waters in South Asia: preventing conflicts and promoting cooperation. Hydrological Sciences Journal, 56 (4), 641-661. Uy, N., Y. Takeuchi, and R. Shaw, 2011: Local adaptation for livelihood resilience in Albay, Philippines. Environmental Hazards, 10 (2), 139-153. Vörösmarty, C.J., P.B. McIntyre, M.O. Gessner, D. Dudgeon, A. Prusevich, P. Green, S. Glidden, S.E. Bunn, C.A. Sullivan, C.R. Liermann, and P.M. Davies, 2010: Global threats to human water security and river biodiversity. Nature, 467 (7315), 555-561. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 van Aalst, M.K., T. Cannon, and I. Burton, 2008: Community level adaptation to climate change: The potential role of participatory community risk assessment. Global Environmental Change-Human and Policy Dimensions, 18 (1), 165-179. van Hooidonk, R., J.A. Maynard, and S. Planes, 2013: Temporary refugia for coral reefs in a warming world. Nature Climate Change, 3 (5), 508-511. Vargas-Silva, C., S. Jha, and G. Sugiyarto, 2009: Remittances in Asia: Implications for the fight against poverty and the pursuit of economic growth. ADB Economics Working Paper Series No. 182, Asian Development Bank, Manila, Philippines, 28 pp. Verchot, L.V., M. Noordwijk, S. Kandji, T. Tomich, C. Ong, A. Albrecht, J. Mackensen, C. Bantilan, K.V. Anupama, and C. Palm, 2007: Climate change: linking adaptation and mitigation through agroforestry. Mitigation and Adaptation Strategies for Global Change, 12 (5), 901-918. Villanoy, C., L. David, O. Cabrera, M. Atrigenio, F. Siringan, P. Alino, and M. Villaluz, 2012: Coral reef ecosystems protect shore from high-energy waves under climate change scenarios. Climatic Change, 112 (2), 493-505. Waddell, S., 2005: Societal learning and change: How governments, business and civil society are creating solutions to complex multi-stakeholder problems. Greenleaf Publishing Ltd., Sheffield, UK, 164 pp. Waddell, S. and S. Khagram, 2007: Multi-stakeholder global networks: emerging systems for the global common good. Partnerships, Governance and Sustainable Development: Reflections on Theory and Practice, 261-287. Wang, G., W. Bai, N. Li, and H. Hu, 2011: Climate changes and its impact on tundra ecosystem in Qinghai-Tibet Plateau, China. Climatic Change, 106 (3), 463-482. Wang, H., 2013: A multi-model assessment of climate change impacts on the distribution and productivity of ecosystems in China. Regional Environmental Change. Wang, X. and H. Liu, 2012: Dynamics change of Betula ermanii population related to shrub and grass on treeline of northern slope of Changbai Mountains. Acta Ecologica Sinica, 32 (10), 3077-3086. Wangdi, K., P. Singhasivanon, T. Silawan, S. Lawpoolsri, N.J. White, and J. Kaewkungwal, 2010: Development of temporal modelling for forecasting and prediction of malaria infections using time-series and ARIMAX analyses: A case study in endemic districts of Bhutan. Malaria Journal, 9, 251. Warner, K., 2010: Global environmental change and migration: Governance challenges. Global Environmental Change, 20 (3), 402-413. Warraich, H., A.K. Zaidi, and K. Patel, 2011: Floods in Pakistan: a public health crisis. Bulletin of the World Health Organization, 89 (3), 236-237. Wassmann, R., S.V.K. Jagadish, S. Heuer, A. Ismail, E. Redona, R. Serraj, R.K. Singh, G. Howell, H. Pathak, and K. Sumfleth, 2009a: Climate change affecting rice production: The physiological and agronomic basis for possible adaptation strategies. In: Advances in Agronomy [Donald, L.S. (ed.)]. Academic Press, pp. 59-122. Wassmann, R., S.V.K. Jagadish, K. Sumfleth, H. Pathak, G. Howell, A. Ismail, R. Serraj, E. Redona, R.K. Singh, and S. Heuer, 2009b: Regional Vulnerability of Climate Change Impacts on Asian Rice Production and Scope for Adaptation. In: Advances in Agronomy [Donald, L.S. (ed.)]. Academic Press, pp. 91-133. WCD (World Commission on Dams), 2000: Dams and Development: A New Framework for Decision-making: the Report of the World Commission on Dams. Earthscan Publication Ltd, 335 pp. Wei, Z., H.J. Jin, J.M. Zhang, S.P. Yu, X.J. Han, Y.J. Ji, R.X. He, and X.L. Chang, 2011: Prediction of permafrost changes in Northeastern China under a changing climate. Science China-Earth Sciences, 54 (6), 924-935. Wilder-Smith, A., Renhorn, K. E., Tissera, H., Abu Bakar, S., Alphey, L., Kittayapong, P., Lindsay, S., Logan, J., Hatz, C., Reiter, P., Rocklo, J., Byass, P., Louis, V. R., Tozan, Y., Massad, E., Tenorio, A., Lagneau, C., Ambert, G. L., Brooks, D., Wegerdt, J. and Gubler, D 2012: Dengue tools: innovative tools and strategies for the surveillance and control of dengue. Global Health Action, 5: 17273. Wilkinson P, S.K., Davies M, Adair H, Armstrong BG, Barrett M, Bruce N, Haines A, Hamilton I, Oreszczyn T, Ridley I, Tonne C, and Chalabi Z., 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: household energy. Lancet, 374 (9705), 1917-1929. Williams, J.W., S.T. Jackson, and J.E. Kutzbacht, 2007: Projected distributions of novel and disappearing climates by 2100 Ad. . Proceedings of the National Academy of Sciences of the United States of America, 104 (14), 5738-5742. Winkel, L.H.E., T.K.T. Pham, M.L. Vi, C. Stengel, M. Amini, T.H. Nguyen, H.V. Pham, and M. Berg, 2011: Arsenic pollution of groundwater in Vietnam exacerbated by deep aquifer exploitation for more than a century. Proceedings of the National Academy of Sciences of the United States of America, 108 (4), 1246-1251. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wong, C.M., T.Q. Thach, P.Y. Chau, E.K. Chan, R.Y. Chung, C.Q. Ou, L. Yang, J.S. Peiris, G.N. Thomas, T.H. Lam, T.W. Wong, and A.J. Hedley, 2010: Part 4. Interaction between air pollution and respiratory viruses: time- series study of daily mortality and hospital admissions in Hong Kong. Research Report of the Health Effects Institute, 154, 283-362. Woodcock J, T.C. Edwards P, Armstrong BG, Ashiru O, Banister D, Beevers S, C.Z. Chalabi Z, Cohen A, Franco OH, Haines A, Hickman R, Lindsay G, and M.D. Mittal I, Tiwari G, Woodward A, Roberts I., 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: urban land transport. Lancet, 374 (9705), 1930-1943. Woodward, F.I. and M.R. Lomas, 2004: Vegetation dynamics - simulating responses to climatic change. Biological Reviews, 79 (3), 643-670. World Bank, 2007: Agriculture for Development: World Development Report 2008. World Bank, Washington D.C, USA, 386 pp. World Bank, 2008: World Databank. Poverty and Inequality Database. [Accessed 18.03.2009]. World Bank, 2010: World Development Report 2010: Development and Climate Change. World Bank. World Bank, 2011: World Development Indicators 2011. World Bank, Washington DC, USA, 460 pp. World Bank, 2013: World Development Indicators Database: DGP per capita 2012. http://data.worldbank.org/indicator/NY.GDP.PCAP.CD [Accessed 27.08.2013]. Wu, Q.B. and T.J. Zhang, 2010: Changes in active layer thickness over the Qinghai-Tibetan Plateau from 1995 to 2007. Journal of Geophysical Research-Atmospheres, 115 (D9), 12 pp. Wu, X., H. Liu, D. Guo, O.A. Anenkhonov, N.K. Badmaeva, and D.V. Sandanov, 2012: Growth decline linked to warming-induced water limitation in hemi-boreal forests. PloS One, 7 (8), 12 pp. Wu, X.H., S.Q. Zhang, X.J. Xu, Y.X. Huang, P. Steinmann, J. Utzinger, T.P. Wang, J. Xu, J. Zheng, and X.N. Zhou, 2008: Effect of floods on the transmission of schistosomiasis in the Yangtze River valley, People's Republic of China. Parasitology International, 57 (3), 271-276. Wyatt, A.B. and I.G. Baird, 2007: Transboundary impact assessment in the Sesan River Basin: The case of the Yali Falls dam. Water Resources Development, 23 (3), 427-442. Xiong, W., I. Holman, E. Lin, D. Conway, J. Jiang, Y. Xu, and Y. Li, 2010: Climate change, water availability and future cereal production in China. Agriculture, Ecosystems & Environment, 135 (1 2), 58-69. Xu, G., T. Chen, X. Liu, L. Jin, W. An, and W. Wang, 2011: Summer temperature variations recorded in tree-ring 13C values on the northeastern Tibetan Plateau. Theoretical and Applied Climatology, 105 (1-2), 51-63. Xu, J., R.E. Grumbine, A. Shrestha, M. Eriksson, X. Yang, Y. Wang, and A. Wilkes, 2009: The melting Himalayas: cascading effects of climate change on water, biodiversity, and livelihoods. Conservation Biology, 23 (3), 520- 530. Xu, K., J.D. Milliman, and H. Xu, 2010: Temporal trend of precipitation and runoff in major Chinese Rivers since 1951. Global and Planetary Change, 73 (3-4), 219-232. Xu, L., R.B. Myneni, F.S. Chapin, III, T.V. Callaghan, J.E. Pinzon, C.J. Tucker, Z. Zhu, J. Bi, P. Ciais, H. Tommervik, E.S. Euskirchen, B.C. Forbes, S.L. Piao, B.T. Anderson, S. Ganguly, R.R. Nemani, S.J. Goetz, P.S.A. Beck, A.G. Bunn, C. Cao, and J.C. Stroeve, 2013: Temperature and vegetation seasonality diminishment over northern lands. Nature Climate Change, 3 (6), 581-586. Xu, X., S. Piao, X. Wang, A. Chen, P. Ciais, and R.B. Myneni, 2012: Spatio-temporal patterns of the area experiencing negative vegetation growth anomalies in China over the last three decades. Environmental Research Letters, 7 (3), 10 pp. Yamanaka, T., Y. Wakiyama, and K. Suzuki, 2012: Is snowmelt runoff timing in the Japanese Alps region shifting toward earlier in the year? Hydrological Research Letters, 7 (1), 1-5. Yamano, H., K. Sugihara, and K. Nomura, 2011: Rapid poleward range expansion of tropical reef corals in response to rising sea surface temperatures. Geophysical Research Letters, 38 (4), 6 pp. Yang, B., C. Qin, K. Huang, Z.X. Fan, and J.J. Liu, 2010: Spatial and temporal patterns of variations in tree growth over the northeastern Tibetan Plateau during the period AD 1450-2001. Holocene, 20 (8), 1235-1245. Yang, Z.-p., J.-x. Gao, L. Zhao, X.-l. Xu, and H. Ouyang, 2013: Linking thaw depth with soil moisture and plant community composition: effects of permafrost degradation on alpine ecosystems on the Qinghai-Tibet Plateau. Plant and Soil, 367 (1-2), 687-700. Yao, T., L. Thompson, W. Yang, W. Yu, Y. Gao, X. Guo, X. Yang, K. Duan, H. Zhao, B. Xu, J. Pu, A. Lu, Y. Xiang, D.B. Kattel, and D. Joswiak, 2012a: Different glacier status with atmospheric circulations in Tibetan Plateau and surroundings. Nature Climate Change, 2 (9), 663-667. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Yao, X.-j., S.-y. Liu, W.-q. Guo, B.-j. Huai, M.-p. Sun, and J.-l. Xu, 2012b: Glacier Change of Altay Mountain in China from 1960 to 2009 -Based on the Second Glacier Inventory of China. Journal of Natural Resources, 27 (10), 1734-1745 [in Chinese]. Yara, Y., M. Vogt, M. Fujii, H. Yamano, C. Hauri, M. Steinacher, N. Gruber, and Y. Yamanaka, 2012: Ocean acidification limits temperature-induced poleward expansion of coral habitats around Japan. Biogeosciences, 9 (12), 4955-4968. Yi, L., W. Jiao, X. Chen, and W. Chen, 2011: An overview of reclaimed water reuse in China. Journal of Environmental Sciences, 23 (10), 1585-1593. Yi, O., Y.-C. Hong, and H. Kim, 2010: Seasonal effect of PM10 concentrations on mortality and morbidity in Seoul, Korea: A temperature-matched case-crossover analysis. Environmental Research, 110 (1), 89-95. Yu, D., M. Chen, Z. Zhou, R. Eric, Q. Tang, and H. Liu, 2013c: Global climate change will severely decrease potential distribution of the East Asian coldwater fish Rhynchocypris oxycephalus (Actinopterygii, Cyprinidae). Hydrobiologia, 700 (1), 23-32. Yu, H., E. Luedeling, and J. Xu, 2010: Winter and spring warming result in delayed spring phenology on the Tibetan Plateau. Proceedings of the National Academy of Sciences of the United States of America, 107 (51), 22151-22156. Yu, H., J. Xu, E. Okuto, and E. Luedeling, 2012: Seasonal response of grasslands to climate change on the Tibetan Plateau. PloS One, 7 (11), 15 pp. Yu, Z., S. Liu, J. Wang, P. Sun, W. Liu, and D.S. Hartley, 2013a: Effects of seasonal snow on the growing season of temperate vegetation in China. Global Change Biology, 19 (7), 2182-2195. Yu, Z., P. Sun, S. Liu, J. Wang, and A. Everman, 2013b: Sensitivity of large-scale vegetation greenup and dormancy dates to climate change in the north-south transect of eastern China. International Journal of Remote Sensing, 34 (20), 7312-7328. Zeller, D., S. Booth, E. Pakhomov, W. Swartz, and D. Pauly, 2011: Arctic fisheries catches in Russia, USA, and Canada: baselines for neglected ecosystems. Polar Biology, 34 (7), 955-973. Zeng, X.D., X.B. Zeng, and M. Barlage, 2008: Growing temperate shrubs over arid and semiarid regions in the Community Land Model-Dynamic Global Vegetation Model. Global Biogeochemical Cycles, 22 (3), 14 pp. Zhang, G., Y. Zhang, J. Dong, and X. Xiao, 2013b: Green-up dates in the Tibetan Plateau have continuously advanced from 1982 to 2011. Proceedings of the National Academy of Sciences of the United States of America, 110 (11), 4309-4314. Zhang, L., H. Guo, L. Ji, L. Lei, C. Wang, D. Yan, B. Li, and J. Li, 2013a: Vegetation greenness trend (2000 to 2009) and the climate controls in the Qinghai-Tibetan Plateau. Journal of Applied Remote Sensing, 7, 18 pp. Zhang, N., T. Yasunari, and T. Ohta, 2011: Dynamics of the larch taiga-permafrost coupled system in Siberia under climate change. Environmental Research Letters, 6 (2). Zhang, T., J. Zhu, and R. Wassmann, 2010: Responses of rice yields to recent climate change in China: An empirical assessment based on long-term observations at different spatial scales (1981 2005). Agricultural and Forest Meteorology, 150 (7 8), 1128-1137. Zhang, Y., B. Peng, and J.E. Hiller, 2008: Weather and the transmission of bacillary dysentery in Jinan, northern China: a time-series analysis. Public Health Reports, 123 (1), 61. Zhao, L., Q.B. Wu, S.S. Marchenko, and N. Sharkhuu, 2010: Thermal state of permafrost and active layer in Central Asia during the International Polar Year. Permafrost and Periglacial Processes, 21 (2), 198-207. Zheng, M., 2011: Resources and eco-environmental protection of salt lakes in China. Environmental Earth Sciences, 64 (6), 1537-1546. Zhou, J., X. Zhang, M. Chen, X. Huang, A. Liu, T. Yang, and H. Tan, 2011: Epidemiological study on hemorrhagic fever with renal syndrome in flood areas. Journal of Central South University, 36 (3), 223. Zhou, S.S., F. Huang, J. Wang, S. Zhang, Y. Su, and L. Tang, 2010: Geographical, meteorological and vectorial factors related to malaria re-emergence in Huang-Huai River of central China. Malaria Journal, 9, 337. Zhou, X.-N., G.-J. Yang, K. Yang, X.-H. Wang, Q.-B. Hong, L.-P. Sun, J.B. Malone, T.K. Kristensen, N.R. Bergquist, and J. Utzinger, 2008: Potential impact of climate change on schistosomiasis transmission in China. The American Journal of Tropical Medicine and Hygiene, 78 (2), 188-194. Ziv, G., E. Baran, S. Nam, I. Rodríguez-Iturbe, and S.A. Levin, 2012: Trading-off fish biodiversity, food security, and hydropower in the Mekong River Basin. Proceedings of the National Academy of Sciences of the United States of America, 109 (15), 5609-5614. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Zuidema, P.A., P.J. Baker, P. Groenendijk, P. Schippers, P. van der Sleen, M. Vlam, and F. Sterck, 2013: Tropical forests and global change: filling knowledge gaps. Trends in Plant Science 18 (8), 413-419. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 24-1: Key risks from climate change and the potential for risk reduction through mitigation and adaptation in Asia. Key risks are identified based on assessment of the literature and expert judgments, with supporting evaluation of evidence and agreement in the referenced chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels are presented for the near-term era of committed climate change (here, for 2030-2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the longer-term era of climate options (here, for 2080-2100), for global mean temperature increase of 2°C and 4°C above preindustrial levels. For each timeframe, risk levels are estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 24-1 (continued) Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 24-2: The amount of information supporting conclusions regarding observed and projected impacts in Asia. / = Relatively abundant/sufficient information; knowledge gaps need to be addressed but conclusions can be drawn based on existing information X = Limited information/no data; critical knowledge gaps, difficult to draw conclusions NR = Not relevant Region Topics/Issues North East Asia Southeast South Central West Asia Asia Asia Asia Asia O= Observed impacts; P = Projected Impacts O P O P O P O P O P O P Freshwater Major river runoff / x / / / / / x x x x x Resources Water supply x x x x x x x x x x x x Terrestrial and Phenology and growth rates / / / / x x x x x x x x Inland Water Distributions of species and biomes / / / / x x x / x x x x Systems Permafrost / / / / / x / / / / / x Inland waters x x / x x x x x x x x x Coastal Coral reefs NR NR / / / / / / NR NR / / Systems and Other coastal ecosystems x x / / x x x x NR NR x x Low-Lying Arctic coast erosion / / NR NR NR NR NR NR NR NR NR NR Areas Food Rice yield x x / / x / x / x x X / Production Wheat yield x x x x x x x / x x / / Systems and Corn yield x x x / x x x x x x x x Food Security Other crops (e.g. barley, potato) x x / / x x x x x X / / Vegetables x x / x x x x x x x x x Fruits x x / x x x x x x x x x Livestock x x / x x x x x x x x x Fisheries and aquaculture production x / x / x / x x x x x x Farming area x / x / x x x / x / x x Water demand for irrigation x / x / x x x / x x x x Pest and disease occurrence x x x x x x x / x x x x Human Floodplains x x / / / / / / x x x x Settlements, Coastal areas x x / / / / / / NR NR x x Industry, and Population and assets x x / / / / / / x x x x Infrastructure Industry and infrastructure x x / / / / / / x x x x Human Health, Health effects of floods x x x x x x / x x x x x Security, Health effects of heat x x / x x x x x x x x x Livelihoods and Health effects of drought x x x x x x x x x x x x Poverty Water-borne diseases x x x x / x / x x x x x Vector-borne diseases x x x x / x / x x x x x Livelihoods and poverty x x / x x x / x x x x x Economic valuation x x x x / / / / x x x x Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 24-1: The land and territories of 51 countries in Asia. NOTE: Currently in production and will be brought to specification using the current UN-accepted maps. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 24-2: Observed and projected changes in annual average temperature and precipitation in Asia. (Top panel, left) Observed temperature trends from 1901-2012 determined by linear regression. [WGI AR5 Figures SPM.1 and 2.21] (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression. [WGI AR5 Figure SPM.2] For observed temperature and precipitation, trends have been calculated where sufficient data permits a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level. Diagonal lines indicate areas where change is not significant. (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and >90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in >66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5. [Boxes 21-3 and CC-RC] Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 24-3: Losses of glacier area in the Altai-Sayan, Pamir and Tien Shan. Remote sensing data analysis from 1960s (Corona) through 2008 (Landsat, ASTER and Alos Prism). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 24 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure TC-1: The intersection of inland and storm surge flooding. Red shows May 5, 2008 MODIS mapping of the tropical cyclone Nargis storm surge along the Irrawaddy Delta and to the east, Myanmar. The blue areas to the north were flooded by the river in prior years. Source: Brakenridge et al., 2013. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 25. Australasia Coordinating Lead Authors Andy Reisinger (New Zealand), Roger Kitching (Australia) Lead Authors Francis Chiew (Australia), Lesley Hughes (Australia), Paul Newton (New Zealand), Sandra Schuster (Australia), Andrew Tait (New Zealand), Penny Whetton (Australia) Contributing Authors Jon Barnett (Australia), Susanne Becken (New Zealand), Paula Blackett (New Zealand), Sarah Boulter (Australia), Andrew Campbell (Australia), Daniel Collins (New Zealand), Jocelyn Davies (Australia), Keith Dear (Australia), Stephen Dovers (Australia), Kyla Finlay (Australia), Bruce Glavovic (New Zealand), Donna Green (Australia), Don Gunasekera (Australia), Simon Hales (New Zealand), John Handmer (Australia), Garth Harmsworth (New Zealand), Alistair Hobday (Australia), Mark Howden (Australia), Graeme Hugo (Australia), David Jones (Australia), Sue Jackson (Australia), Darren King (New Zealand), Miko Kirschbaum (New Zealand), Jo Luck (Australia), Jan McDonald (Australia), Kathy McInnes (Australia), Yiheyis Maru (Australia), Johanna Mustelin (Australia), Barbara Norman (Australia), Grant Pearce (New Zealand), Susan Peoples (New Zealand), Ben Preston (USA), Joseph Reser (Australia), Penny Reyenga (Australia), Mark Stafford-Smith (Australia), Xiaoming Wang (Australia), Leanne Webb (Australia) Review Editors Blair Fitzharris (New Zealand), David Karoly (Australia) Contents Executive Summary 25.1. Introduction and Major Conclusions from Previous Assessments 25.2. Observed and Projected Climate Change 25.3. Socio-Economic Trends Influencing Vulnerability and Adaptive Capacity 25.3.1. Economic, Demographic and Social Trends 25.3.2. Use and Relevance of Socio-Economic Scenarios in Adaptive Capacity/Vulnerability Assessments 25.4. Cross-Sectoral Adaptation: Approaches, Effectiveness, and Constraints 25.4.1. Frameworks, Governance, and Institutional Arrangements 25.4.2. Constraints on Adaptation and Leading Practice Models 25.4.3. Socio-cultural Factors Influencing Impacts of and Adaptation to Climate Change 25.5. Freshwater Resources 25.5.1. Observed Impacts 25.5.2. Projected Impacts 25.5.3. Adaptation 25.6. Natural Ecosystems 25.6.1. Inland Freshwater and Terrestrial Ecosystems 25.6.1.1. Observed Impacts 25.6.1.2. Projected Impacts 25.6.1.3. Adaptation 25.6.2. Coastal and Ocean Ecosystems 25.6.2.1. Observed Impacts Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 25.6.2.2. Projected Impacts 25.6.2.3. Adaptation 25.7. Major Industries 25.7.1. Production Forestry 25.7.1.1. Observed and Projected Impacts 25.7.1.2. Adaptation 25.7.2. Agriculture 25.7.2.1. Projected Impacts and Adaptation Livestock Systems 25.7.2.2. Projected Impacts and Adaptation Cropping 25.7.2.3. Integrated Adaptation Perspectives 25.7.3. Mining 25.7.4. Energy Supply, Transmission, and Demand 25.7.5. Tourism 25.7.5.1. Projected Impacts 25.7.5.2. Adaptation 25.8. Human Society 25.8.1. Human Health 25.8.1.1. Observed Impacts 25.8.2.2. Projected Impacts 25.8.3.3. Adaptation 25.8.2. Indigenous Peoples 25.8.2.1. Aboriginal and Torres Strait Islanders 25.8.2.2. New Zealand M ori 25.9. Interactions among Impacts, Adaptation, and Mitigation Responses 25.9.1. Interactions among Local-Level Impacts, Adaptation, and Mitigation Responses 25.9.2. Intra- and Inter-Regional Flow-On Effects between Impacts, Adaptation, and Mitigation 25.10. Synthesis and Regional Key Risks 25.10.1. Economy-wide Impacts and Potential of Mitigation to Reduce Risks 25.10.2. Regional Key Risks as a Function of Mitigation and Adaptation 25.10.3. The Role of Adaptation in Managing Key Risks, and Adaptation limits 25.11. Filling Knowledge Gaps to Improve Management of Climate Risks References Chapter Boxes 25-1. Coastal Adaptation Planning and Legal Dimensions 25-2. Adaptation through Water Resources Policy and Management in Australia 25-3. Impacts of a Changing Climate in Natural and Managed Ecosystems 25-4. Biosecurity 25-5. Climate Change Vulnerability and Adaptation in Rural Areas 25-6. Climate Change and Fire 25-7. Insurance as Climate Risk Management Tool 25-8. Changes in Flood Risk and Management Responses 25-9. Opportunities, Constraints, and Challenges to Adaptation in Urban Areas 25-10. Land-based Interactions Among Climate, Energy, Water, and Biodiversity Frequently Asked Questions 25.1: How can we adapt to climate change if projected future changes remain uncertain? 25.2: What are the key risks from climate change to Australia and New Zealand? Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Executive Summary The regional climate is changing (very high confidence). The region continues to demonstrate long term trends toward higher surface air and sea-surface temperatures, more hot extremes and fewer cold extremes, and changed rainfall patterns. Over the past 50 years, increasing greenhouse gas concentrations have contributed to rising average temperature in Australia (high confidence) and New Zealand (medium confidence) and decreasing rainfall in south- western Australia (high confidence). [25.2, Table 25-1] Warming is projected to continue through the 21st century (virtually certain) along with other changes in climate. Warming is expected to be associated with rising snow lines (very high confidence), more frequent hot extremes, less frequent cold extremes (high confidence), and increasing extreme rainfall related to flood risk in many locations (medium confidence). Annual average rainfall is expected to decrease in south-western Australia (high confidence) and elsewhere in most of far southern Australia and the north-east South Island and northern and eastern North Island of New Zealand (medium confidence), and to increase in other parts of New Zealand (medium confidence). Tropical cyclones are projected to increase in intensity but remain similar or decrease in numbers (low confidence), and fire weather is projected to increase in most of southern Australia (high confidence) and many parts of New Zealand (medium confidence). Regional sea level rise will very likely exceed the historical rate (1971-2010), consistent with global mean trends. [25.2, Table 25-1, Box 25-6; WGI 13.5, 13.6] Uncertainty in projected rainfall changes remains large for many parts of Australia and New Zealand, which creates significant challenges for adaptation. For example, projections for average annual runoff in far south- eastern Australia range from little change to a 40% decline for 2°C global warming above current levels. The dry end of these scenarios would have severe implications for agriculture, rural livelihoods, ecosystems and urban water supply, and would increase the need for transformational adaptation (high confidence). [25.2, 25.5.1, 25.6.1, 25.7.2, Box 25-2, Box 25-5] Recent extreme climatic events show significant vulnerability of some ecosystems and many human systems to current climate variability (very high confidence), and the frequency and/or intensity of such events is projected to increase in many locations (medium to high confidence). For example, high sea surface temperatures have repeatedly bleached coral reefs in north-eastern Australia (since the late 1970s) and more recently in western Australia. Recent floods in Australia and New Zealand caused severe damage to infrastructure and settlements and 35 deaths in Queensland alone (2011); the Victorian heat wave (2009) increased heat-related morbidity and was associated with more than 300 excess deaths, while intense bushfires destroyed over 2,000 buildings and led to 173 deaths; and widespread drought in south-east Australia (1997-2009) and many parts of New Zealand (2007-2009; 2012-13) resulted in substantial economic losses (e.g. regional GDP in the southern Murray Darling Basin was below forecast by about 5.7% in 2007/08, and New Zealand lost about NZ$3.6b in direct and off-farm output in 2007-09). [Table 25-1, 25.6.2, 25.8.1, Box 25-5, Box 25-6, Box 25-8] Without adaptation, further changes in climate, atmospheric CO2 and ocean acidity are projected to have substantial impacts on water resources, coastal ecosystems, infrastructure, health, agriculture and biodiversity (high confidence). Freshwater resources are projected to decline in far south-west and far south-east mainland Australia (high confidence) and for rivers originating in the north-east of the South Island and east and north of the North Island of New Zealand (medium confidence). Rising sea levels and increasing heavy rainfall are projected to increase erosion and inundation, with consequent damages to many low-lying ecosystems, infrastructure and housing; increasing heat waves will increase risks to human health; rainfall changes and rising temperatures will shift agricultural production zones; and many native species will suffer from range contractions and some may face local or even global extinction. [25.5.1, 25.6.1, 25.6.2, 25.7.2, 25.7.4, Box 25-1, Box 25-5, Box 25-8] Some sectors in some locations have the potential to benefit from projected changes in climate and increasing atmospheric CO2 (high confidence). Examples include reduced winter mortality (low confidence), reduced energy demand for winter heating in New Zealand and southern parts of Australia, and forest growth in cooler regions Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 except where soil nutrients or rainfall are limiting. Spring pasture growth in cooler regions would also increase and be beneficial for animal production if it can be utilized. [25.7.1, 25.7.2, 25.7.4, 25.8.1] Adaptation is already occurring and adaptation planning is becoming embedded in some planning processes, albeit mostly at the conceptual rather than implementation level (high confidence). Many solutions for reducing energy and water consumption in urban areas with co-benefits for climate change adaptation (e.g. greening cities and recycling water) are already being implemented. Planning for reduced water availability in southern Australia and for sea-level rise in both countries is becoming adopted widely, although implementation of specific policies remains piecemeal, subject to political changes, and open to legal challenges. [25.4, Box 25-1, Box 25-2, Box 25-9] Adaptive capacity is generally high in many human systems, but implementation faces major constraints especially for transformational responses at local and community levels (high confidence). Efforts to understand and enhance adaptive capacity and adaptation processes have increased since the AR4, particularly in Australia. Constraints on implementation arise from: absence of a consistent information base and uncertainty about projected impacts; limited financial and human resources to assess local risks and to develop and implement effective policies and rules; limited integration of different levels of governance; lack of binding guidance on principles and priorities; different attitudes towards the risks associated with climate change, and different values placed on objects and places at risk. [25.4, 25.10.3, Box 25-1, Table 25-2] Indigenous peoples in both Australia and New Zealand have higher than average exposure to climate change due to a heavy reliance on climate-sensitive primary industries and strong social connections to the natural environment, and face particular constraints to adaptation (medium confidence). Social status and representation, health, infrastructure and economic issues, and engagement with natural resource industries constrain adaptation and are only partly offset by intrinsic adaptive capacity (high confidence). Some proposed responses to climate change may provide economic opportunities, particularly in New Zealand related to forestry. Torres Strait communities are vulnerable even to small sea level rises (high confidence). [25.3, 25.8.2] We identify eight regional key risks during the 21st century based on the severity of potential impacts for different levels of warming, uniqueness of the systems affected, and adaptation options (high confidence). These risks differ in the degree to which they can be managed via adaptation and mitigation, and some are more likely to be realized than others, but all warrant attention from a risk-management perspective. Some potential impacts can be delayed but now appear very difficult to avoid entirely, even with globally effective mitigation and planned adaptation: o significant change in community composition and structure of coral reef systems in Australia, driven by increasing sea-surface temperatures and ocean acidification; the ability of corals to adapt naturally to rising temperatures and acidification appears limited and insufficient to offset the detrimental effects [25.6.2, 30.5, Box CC-CR] o loss of montane ecosystems and some native species in Australia, driven by rising temperatures and snow lines, increased fire risk and drying trends; fragmentation of landscapes, limited dispersal and limited rate of evolutionary change constrain adaptation options [25.6.1] Some impacts have the potential to be severe but can be reduced substantially by globally effective mitigation combined with adaptation, with the need for transformational adaptation increasing with the rate and magnitude of climate change: o increased frequency and intensity of flood damage to settlements and infrastructure in Australia and New Zealand, driven by increasing extreme rainfall although the amount of change remains uncertain; in many locations, continued reliance on increased protection alone would become progressively less feasible [Table 25-1, 25.4.2, Box 25-8, 25.10.3] o constraints on water resources in southern Australia, driven by rising temperatures and reduced cool- season rainfall; integrated responses encompassing management of supply, recycling, water conservation and increased efficiency across all sectors are available and some are being implemented in areas already facing shortages [25.2, 25.5.2, Box 25-2] o increased morbidity, mortality and infrastructure damages during heat waves in Australia, resulting from increased frequency and magnitude of extreme high temperatures; vulnerable populations include Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 the elderly and those with existing chronic diseases; population increases and ageing trends constrain effectiveness of adaptation responses [25.8.1] o increased damages to ecosystems and settlements, economic losses and risks to human life from wildfires in most of southern Australia and many parts of New Zealand, driven by rising temperatures and drying trends; local planning mechanisms, building design, early warning systems and public education can assist with adaptation and are being implemented in regions that have experienced major events [25.2, Table 25-1, 25.6.1, 25.7.1, Box 25-6] For some impacts, severity depends on changes in climate variables that span a particularly large range, even for a given global temperature change. The most severe changes would present major challenges if realized: o increasing risks to coastal infrastructure and low-lying ecosystems in Australia and New Zealand from continuing sea level rise, with widespread damages towards the upper end of projected changes; managed retreat is a long-term adaptation strategy for human systems but options for some natural ecosystems are limited due to the rapidity of change and lack of suitable space for landward migration. Risks from sea level rise continue to increase beyond 2100 even if temperatures are stabilised. [Table 25-1, 25.4.2, Box 25-1, 25.6.1, 25.6.2; WGI 13.5] o significant reduction in agricultural production in the Murray-Darling Basin and far south-eastern and south-western Australia if scenarios of severe drying are realised; more efficient water use, allocation and trading would increase the resilience of systems in the near term but cannot prevent significant reductions in agricultural production and severe consequences for ecosystems and some rural communities at the dry end of the projected changes [25.2, 25.5.2, 25.7.2, Box 25-2, Box 25-5] Significant synergies and trade-offs exist between alternative adaptation responses, and between mitigation and adaptation responses; interactions occur both within Australasia and between Australasia and the rest of the world (very high confidence). Increasing efforts to mitigate and adapt to climate change imply an increasing complexity of interactions, particularly at the intersections among water, energy and biodiversity, but tools to understand and manage these interactions remain limited. Flow-on effects from climate change impacts and responses outside Australasia have the potential to outweigh some of the direct impacts within the region, particularly economic impacts on trade-intensive sectors such as agriculture (medium confidence) and tourism (high agreement, limited evidence), but they remain among the least explored issues. [25.7.5, 25.9.1, 25.9.2, Box 25-10] Understanding of future vulnerability of human and mixed human-natural systems to climate change remains limited due to incomplete consideration of socio-economic dimensions (very high confidence). Future vulnerability will depend on factors such as wealth and its distribution across society, patterns of ageing, access to technology and information, labour force participation, societal values, and mechanisms and institutions to resolve conflicts. These dimensions have received only limited attention and are rarely included in vulnerability assessments, and frameworks to integrate social, psychological and cultural dimensions of vulnerability with bio- physical impacts and economic losses are lacking. In addition, conclusions for New Zealand in many sectors, even for bio-physical impacts, are based on limited studies that often use a narrow set of assumptions, models and data and hence have not explored the full range of potential outcomes. [25.3, 25.4, 25.11] 25.1. Introduction and Major Conclusions from Previous Assessments Australasia is defined here as lands, territories, offshore waters and oceanic islands of the exclusive economic zones of Australia and New Zealand. Both countries are relatively wealthy with export-led economies. Both have Westminster-style political systems and have a relatively recent history of non-indigenous settlement (Australia in the late 18th, New Zealand in the early 19th century). Both retain significant indigenous populations. Principal findings from the IPCC Fourth Assessment Report (AR4) for the region were (Hennessy et al., 2007): Consistent with global trends, Australia and New Zealand had experienced warming of 0.4 to 0.7°C since 1950 with changed rainfall patterns and sea-level rise of about 70 mm across the region; there had also been a greater frequency and intensity of droughts and heat waves, reduced seasonal snow cover and glacial retreat. Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Impacts from recent climate changes were evident in increasing stresses on water supply and agriculture, and changed natural ecosystems; some adaptation had occurred in these sectors but vulnerability to extreme events such as fire, tropical cyclones, droughts, hail and floods remained high. The climate of the 21st century would be warmer (virtually certain), with changes in extreme events including more intense and frequent heat waves, fire, floods, storm surges and droughts but less frequent frost and snow (high confidence), reduced soil moisture in large parts of the Australian mainland and eastern New Zealand but more rain in western New Zealand (medium confidence). Significant advances had occurred in understanding future impacts on water, ecosystems, Indigenous people and health together with an increased focus on adaptation; potential impacts would be substantial without further adaptation, particularly for water security, coastal development, biodiversity, and major infrastructure, but impacts on agriculture and forestry would be variable across the region, including potential benefits in some areas. Vulnerability would increase mainly due to an increase in extreme events; human systems were considered to have a higher adaptive capacity than natural systems. Hotspots of high vulnerability by 2050 under a medium emissions scenario included: - significant loss of biodiversity in areas such as alpine regions, the Wet Tropics, the Australian south-west, Kakadu wetlands, coral reefs and sub-Antarctic islands; - water security problems in the Murray-Darling basin, south-western Australia and eastern New Zealand; - potentially large risks to coastal development in south-eastern Queensland and in New Zealand from Northland to the Bay of Plenty. 25.2. Observed and Projected Climate Change Australasia exhibits a wide diversity of climates, such as moist tropical monsoonal, arid and moist temperate, including alpine conditions. Key climatic processes are the Asian-Australian monsoon and the southeast trade winds over northern Australia, and the subtropical high pressure belt and the mid-latitude storm tracks over southern Australia and New Zealand. Tropical cyclones also affect northern Australia, and, more rarely, ex-tropical cyclones affect some parts of New Zealand. Natural climatic variability is very high in the region, especially for rainfall and over Australia, with the El Nino-Southern Oscillation (ENSO) being the most important driver (McBride and Nicholls, 1983; Power et al., 1998; Risbey et al., 2009). The southern annular mode, Indian Ocean Dipole and the Interdecadal Pacific Oscillation are also important regional drivers (Thompson and Wallace, 2000; Salinger et al., 2001; Cai et al., 2009b). This variability poses particular challenges for detecting and projecting anthropogenic climate change and its impacts in the region. For example, changes in ENSO in response to anthropogenic climate change are uncertain (AR5 WGI Ch14) but, given current ENSO impacts, any changes would have the potential to significantly influence rainfall and temperature extremes, droughts, tropical cyclones, marine conditions and glacial mass balance (Mullan, 1995; Chinn et al., 2005; Holbrook et al., 2009; Diamond et al., 2012; Min et al., 2013). Understanding of observed and projected climate change has received much attention since AR4, particularly in Australia, with a focus on the causes of observed rainfall changes and more systematic analysis of projected changes from different models and approaches. Climatic extremes have also been a research focus. Table 25-1 presents an assessment of this body of research for observed trends and projected changes for a range of climatic variables (including extremes) relevant for regional impacts and adaptation, including examples of the magnitude of projected change, and attribution, where possible. Most studies are based on CMIP3 models and SRES scenarios, but CMIP5 model results are considered where available (see also AR5 WGI Chap 14 & Atlas; WGII Chapter 21). The region has exhibited warming to the present (very high confidence) and is virtually certain to continue to do so (Table 25-1). Observed and CMIP5-modelled past and projected future annual average surface temperatures are shown in Figures 25-1 and 25-2. For further details see WGI Atlas, AI.68-69. Changes in precipitation have been observed with very high confidence in some areas over a range of time scales, such as increases in north-western Australia since the 1950s, the autumn/winter decline since 1970 in south-western Australia and, since the 1990s, in south-eastern Australia, and over 1950-2004 increases in annual rainfall in the south and west of the South Island and west of the North Island of New Zealand, and decreases in the north-east of the South Island and east and north of the North Island. Based on multiple lines of evidence, annual average rainfall is projected to decrease with high Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 confidence in south-western Australia. For New Zealand, annual average rainfall is projected to decrease in the north-eastern South Island and eastern and northern North Island, and increase in other parts of the country (medium confidence). The direction and magnitude of rainfall change in eastern and northern Australia remains a key uncertainty (Table 25-1). This pattern of projected rainfall change is reflected in annual average CMIP5 model results (Figure 25-1), but with important additional dimensions relating to seasonal changes and spread across models (see also WGI Atlas, AI.70- 71). Examples of the magnitude of projected annual change from 1990 to 2090 (percent model mean change +/- intermodel standard deviation) under RCP8.5 from CMIP5 are -20+/-13% in south-western Australia, -2+/-21% in the Murray Darling Basin, and -5+/-22% in southeast Queensland (Irving et al., 2012). Projected changes during winter and spring are more pronounced and/or consistent across models than the annual changes, e.g. drying in south- western Australia (-32+/-11%, June to August), the Murray Darling Basin (-16+/-22%, June to August), and southeast Queensland (-15+/-26%, September to November), whereas there are increases of 15% or more in the west and south of the South Island of New Zealand (Irving et al., 2012). Downscaled CMIP3 model projections for New Zealand indicate a stronger drying pattern in the south-east of the South Island and eastern and northern regions of the North Island in winter and spring (Reisinger et al., 2010) than seen in the raw CMIP5 data; based on similar broader scale changes this pattern is expected to hold once CMIP5 data are also downscaled (Irving et al., 2012). Other projected changes of at least high confidence include regional increases in sea surface temperature, the occurrence of hot days, fire weather in southern Australia, mean and extreme sea level, and ocean acidity (see WGI 6.4.4 for projections); and decreases in cold days and snow extent and depth. Although changes to tropical cyclone occurrence and that of other severe storms are potentially important for future vulnerability, regional changes to these phenomena cannot be projected with at least medium confidence as yet (see Table 25-1). [INSERT FIGURE 25-1 HERE Figure 25-1: Observed and projected changes in annual average temperature and precipitation. (Top panel, left) Observed temperature trends from 1901-2012 determined by linear regression [WGI AR5 Figures SPM.1 and 2.21]. (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression [WGI AR5 Figure SPM.2]. For observed temperature and precipitation, trends have been calculated where sufficient data permits a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level. Diagonal lines indicate areas where change is not significant. (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and >90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in >66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5 [Boxes 21-3 and CC-RC].] [INSERT FIGURE 25-2 HERE Figure 25-2: Observed and simulated variations in past and projected future annual average near-surface air temperature over land areas of Australia (left) and New Zealand (right). Black lines show various estimates from observational measurements. Shading denotes the 5-95 percentile range of climate model simulations driven with historical changes in anthropogenic and natural drivers (63 simulations), historical changes in natural drivers only (34), the RCP2.6 emissions scenario (63), and the RCP8.5 (63). Data are anomalies from the 1986-2005 average of the individual observational data (for the observational time series) or of the corresponding historical all- forcing simulations. Further details are given in Box 21-3.] Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT TABLE 25-1 HERE Table 25-1: Observed and projected changes in key climate variables, and (where assessed) the contribution of human activities to observed changes. For further relevant information see WGI Chapters 3, 6 (ocean changes, including acidification), 11, 12 (projections), 13 (sea level) and 14 (regional climate phenomena).] 25.3. Socio-Economic Trends Influencing Vulnerability and Adaptive Capacity 25.3.1. Economic, Demographic and Social Trends The economies of Australia and New Zealand rely on natural resources, agriculture, minerals, manufacturing and tourism, but the relative importance of these sectors differs between the two countries. Agriculture and mineral/energy resources accounted, respectively, for 11% and 55% (Australia) and 56% and 5% (New Zealand) of the value of total exports in 2010/2011 (ABS, 2012c; SNZ, 2012b). Water abstraction per capita in both countries is in the top half of the OECD, decreasing since 1990 in Australia but increasing in New Zealand; more than half is used for irrigation (OECD, 2010, 2013a). Between 1970 and 2011, GDP grew by an average of 3.2% p.a. in Australia and 2.4% p.a. in New Zealand, with annual GDP per capita growth of 1.8% and 1.2%, respectively (SNZ, 2011; ABS, 2012d). GDP is projected to grow on average by 2.5-3.5% p.a. in Australia and about 1.9% p.a. in New Zealand to 2050 (Australian Treasury, 2010; Bell et al., 2010) but subject to significant shorter-term fluctuations. The populations of Australia and New Zealand are projected to grow significantly over at least the next several decades (very high confidence; ABS, 2008; SNZ, 2012a); Australia s population from 22.3 million in 2011 to 31-43 million by 2056 and 34-62 million by 2101 (ABS, 2008, 2013); New Zealand s population from 4.4 million in 2011 to 5.1-7.1 million by 2061 (SNZ, 2012a). The number of people aged 65 and over is projected to almost double in the next two decades (ABS, 2008; SNZ, 2012a). More than 85% of the Australasian population lives in urban areas and their satellite communities, mostly in coastal areas (DCC, 2009; SNZ, 2010b; UN, 2012; see Box 25-9). Urban concentration and depletion of remote rural areas is expected to continue (Mendham and Curtis, 2010; SNZ, 2010c; Box 25-5), but some coastal non-urban spaces also face increasing development pressure (Freeman and Cheyne, 2008; Gurran, 2008; Box 25-1). More than 20% of Australasian residents were born overseas (OECD, 2013a). Poverty rates and income inequality in Australia and New Zealand are in the upper half of OECD countries, and both measures increased significantly in both countries between the mid-1980s and the late-2000s (OECD, 2013a). Measurement of poverty and inequality, however, is highly contested, and it remains difficult to anticipate future changes and their effects on adaptive capacity (Peace, 2001; Scutella et al., 2009; 25.3.2). Indigenous peoples constitute about 2.5% and 15% of the Australian and New Zealand populations, respectively, but in Australia, their national share is growing and they constitute a much higher percentage of the population in remote and very remote regions (ABS, 2009, 2010b; SNZ, 2010a). Indigenous peoples in both countries have lower than average life expectancy, income and education, implying that changes in socio-economic status and social inclusion could strongly influence their future adaptive capacity (see 25.8.2). 25.3.2. Use and Relevance of Socio-Economic Scenarios in Adaptive Capacity/Vulnerability Assessments Demographic, economic and socio-cultural trends influence the vulnerability and adaptive capacity of individuals and communities (see Chapters 2, 11-13, 16, 20). A limited but growing number of studies in Australasia have attempted to incorporate such information, e.g. changes in the number of people and percentage of elderly people at risk (Preston et al., 2008; Baum et al., 2009; Preston and Stafford-Smith, 2009; Roiko et al., 2012), the density of urban settlements and exposed infrastructure (Preston and Jones, 2008; Preston et al., 2008; Baynes et al., 2012), population-driven pressures on water demand (Jollands et al., 2007; CSIRO, 2009), and economic and social factors affecting individual coping, planning and recovery capacity (Dwyer et al., 2004; Khan, 2012; Roiko et al., 2012). Socio-economic considerations are used increasingly to understand adaptive capacity of communities (Preston et al., 2008; Smith et al., 2008; Fitzsimons et al., 2010; Soste, 2010; Brunckhorst et al., 2011) and to construct scenarios to help build regional planning capacity (Energy Futures Forum, 2006; Frame et al., 2007; Pride et al., 2010; Pettit et Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 al., 2011; Taylor et al., 2011). Such scenarios, however, are only beginning to be used to quantify vulnerability to climate change (except e.g. Bohensky et al., 2011; Baynes et al., 2012; Low Choy et al., 2012). Apart from these emerging efforts, most vulnerability studies from Australasia make no or very limited use of socio- economic factors, consider only current conditions, and/or rely on postulated correlations between generic socio- economic indicators and climate change vulnerability. In many cases this limits confidence in conclusions regarding future vulnerability to climate change and adaptive capacity of human and mixed natural-human systems. 25.4. Cross-Sectoral Adaptation: Approaches, Effectiveness, and Constraints 25.4.1. Frameworks, Governance, and Institutional Arrangements Adaptation responses depend heavily on institutional and governance arrangements (see Chapters 2, 14, 15, 16, 20). Responsibility for development and implementation of adaptation policy in Australasia is largely devolved to local governments and, in Australia, to State governments and Natural Resource Management bodies. Federal/central government supports adaptation mostly via provision of information, tools, legislation, policy guidance and (in Australia) support for pilot projects. A standard risk management paradigm has been promoted to embed adaptation into decision-making practices (AGO, 2006; MfE, 2008b; Standards Australia, 2013), but broader systems and resilience approaches are used increasingly for natural resource management (Clayton et al., 2011; NRC, 2012). The Council of Australian Governments agreed a national adaptation policy framework in 2007 (COAG, 2007). This included establishing the collaborative National Climate Change Adaptation Research Facility (NCCARF) in 2008, which complemented CSIRO s Climate Adaptation Flagship. The federal government supported a first-pass national coastal risk assessment (DCC, 2009; DCCEE, 2011), is developing indicators and criteria for assessing adaptation progress and outcomes (DIICCSRT, 2013), and commissioned targeted reports addressing impacts and management options for natural and managed landscapes (Campbell, 2008; Steffen et al., 2009; Dunlop et al., 2012), National and World Heritage areas (ANU, 2009; BMT WBM, 2011), and indigenous and urban communities (Green et al., 2009; Norman, 2010). Most State and Territory governments have also developed adaptation plans (e.g. DSE, 2013). In New Zealand, the central government updated and expanded tools to support impact assessments and adaptation responses consistent with regulatory requirements (MfE, 2008b, c, d, 2010b), and revised key directions for coastal management (Minister of Conservation, 2010). No cross-sectoral adaptation policy framework or national-level risk assessments exist, but some departments commissioned high-level impacts and adaptation assessments after the AR4 (e.g. on agriculture and on biodiversity; Wratt et al., 2008; McGlone and Walker, 2011; Clark et al., 2012). Public and private sector organisations are potentially important adaptation actors but exhibit large differences in preparedness, linked to knowledge about climate change, economic opportunities, external connections, size, and scope for strategic planning (Gardner et al., 2010; Taylor et al., 2012a; Johnston et al., 2013; Kuruppu et al., 2013; see also Chapters 10, 16). This creates challenges for achieving holistic societal outcomes (see also 25.7-25.9). Several recent policy initiatives in Australia, while responding to broader socio-economic and environmental pressures, include goals to reduce vulnerability to climate variability and change. These include establishing the Murray-Darling Basin Authority to address over-allocation of water resources (Connell and Grafton, 2011; MDBA, 2011), removal of the interest rate subsidy during exceptional droughts (Productivity Commission, 2009), and management of bush fire and flood risk (VBRC, 2010; QFCI, 2012). These may be seen as examples of mainstreaming adaptation (Dovers, 2009), but they also demonstrate lag times in policy design and implementation, windows of opportunity presented by crises (e.g. the Millennium Drought 1997-2009, the Victorian bushfires 2009 and Queensland floods of 2011), and the challenges arising from competing interests in managing finite and changing water resources (Botterill and Dovers, 2013; Pittock, 2013; Box 25-2). Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 25.4.2. Constraints on Adaptation and Emerging Leading Practice Models A rapidly growing literature since the AR4 confirms, with high confidence, that while the adaptive capacity of society in Australasia is generally high, there are formidable environmental, economic, informational, social, attitudinal and political constraints, especially for local governments and small or highly fragmented industries. Reviews of public- and private-sector adaptation plans and strategies in Australia demonstrate strong efforts in institutional capacity building, but differences in assessment methods and weaknesses in translating goals into specific policies (White, 2009; Gardner et al., 2010; Measham et al., 2011; Preston et al., 2011; Kay et al., 2014). Similarly, local governments in New Zealand to date have focused mostly on impacts and climate-related hazards; some have developed adaptation plans, but few have committed to specific policies and steps to implementation (e.g. O'Donnell, 2007; Britton, 2010; Fitzharris, 2010; HRC, 2012; KCDC, 2012; Lawrence et al., 2013b). Table 25-2 summarises key constraints and corresponding enabling factors for effective institutional adaptation processes identified in Australia and New Zealand. Scientific uncertainty and resource limitations are reported consistently as important constraints, particularly for smaller councils. Ultimately more powerful constraints arise, however, from current governance and legislative arrangements and the lack of consistent tools to deal with dynamic risks and uncertainty or to evaluate the success of adaptation responses (high agreement, robust evidence; Britton, 2010; Barnett et al., 2013; Lawrence et al., 2013b; Mukheibir et al., 2013; Webb et al., 2013; see also Chapter 16). [INSERT TABLE 25-2 HERE Table 25-2: Constraints and enabling factors for institutional adaptation processes in Australasia.] Some constraints exacerbate others. There is high confidence that the absence of a consistent information base and binding guidelines that clarify governing principles and liabilities is a challenge particularly for small and resource- limited local authorities, which need to balance special interest advocacy with longer term community resilience. This heightens reliance on individual leadership subject to short-term political change and can result in piecemeal and inconsistent risk assessments and responses between levels of government and locations, and over time (Smith et al., 2008; Brown et al., 2009; Norman, 2009; Britton, 2010; Rouse and Norton, 2010; Abel et al., 2011; McDonald, 2011; Rive and Weeks, 2011; Corkhill, 2013; Macintosh et al., 2013). In these situations, planners tend to rely more on single numbers for climate projections that can be argued in court (Reisinger et al., 2011; Lawrence et al., 2013b), which increases the risk of maladaptation given the uncertain and dynamic nature of climate risk (McDonald, 2010; Stafford-Smith et al., 2011; Gorddard et al., 2012; McDonald, 2013; Reisinger et al., 2014). Vulnerability assessments that take mid- to late-century impacts as their starting point can inhibit actors from implementing adaptation actions, as distant impacts are easily discounted and difficult to prioritise in competition with near-term non-climate change pressures (Productivity Commission, 2012). Emerging leading practice models in Australia (Balston, 2012; HCCREMS, 2012; SGS, 2012) and New Zealand (MfE, 2008a; Britton et al., 2011) recommend a high-level scan of sectors and locations at risk and emphasise a focus on near-term decisions that influence current and future vulnerability (which could range from early warning systems to strategic and planning responses). More detailed assessment can then focus on this more tractable subset of issues, based on explicit and iterative framing of the adaptation issue (Webb et al., 2013) and taking into account the full lifetime (lead- and consequence time) of the decision/asset in question (Stafford-Smith et al., 2011). Participatory processes help balance societal preferences with robust scientific information and ensure ownership by affected communities but rely on human capital and political commitment (high confidence; Hobson and Niemeyer, 2011; Rouse and Blackett, 2011; Weber et al., 2011; Leitch and Robinson, 2012). Realising widespread and equitable participation is challenging where policies are complex, debates polarised, legitimacy of institutions contested and potential transformational changes threaten deeply held values (Gardner et al., 2009a; Gorddard et al., 2012; Burton and Mustelin, 2013; see also 25.4.3). Regional approaches that engage diverse stakeholders, government and science providers and support the co-production of knowledge can help overcome some of these problems but require long-term institutional and financial commitments (e.g. Britton et al., 2011; DSEWPC, 2011; CSIRO, 2012; IOCI, 2012; Low Choy et al., 2012; Webb and Beh, 2013). Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 There is active debate about the extent to which incremental adjustments of existing planning instruments, institutions and decision-making processes can deal adequately with the dynamic and uncertain nature of climate change and support transformational responses (Kennedy et al., 2010; Preston et al., 2011; Park et al., 2012; Dovers, 2013; Lawrence et al., 2013b; McDonald, 2013; Stafford-Smith, 2013). Recent studies suggest a greater focus on flexibility and matching decision-making frameworks to specific problems (Hertzler, 2007; Nelson et al., 2008; Dobes, 2010; Howden and Stokes, 2010; Randall et al., 2012). Limitations of mainstreamed and autonomous adaptation and the case for more proactive government intervention are being explored in Australia (Productivity Commission, 2012; Johnston et al., 2013), but have not yet resulted in new policy frameworks. _____ START BOX 25-1 HERE _____ Box 25-1. Coastal Adaptation Planning and Legal Dimensions Sea level rise is a significant risk for Australia and New Zealand (very high confidence) due to intensifying coastal development and the location of population centres and infrastructure (see 25.3). Under a high emissions scenario (RCP8.5), global mean sea level would likely rise by 0.53 to 0.97 m by 2100, relative to 1986-2005, whereas with stringent mitigation (RCP2.6), the likely rise by 2100 would be 0.28 to 0.6 m (medium confidence). Based on current understanding, only instability of the Antarctic Ice Sheet, if initiated, could lead to a rise substantially above the likely range; evidence remains insufficient to evaluate its probability, but there is medium confidence that this additional contribution would not exceed several tenths of a metre during the 21st century (AR5 WGI 13.5). Local case studies in New Zealand (Fitzharris, 2010; Reisinger et al., 2014) and national reviews in Australia (DCC, 2009; DCCEE, 2011) demonstrate risks to large numbers of residential and commercial assets as well as key services, with widespread damages at the upper end of projected ranges (high confidence). In Australia, sea level rise of 1.1 m would affect over A$226 billion of assets, including up to 274,000 residential and 8,600 commercial buildings (DCCEE, 2011), with additional intangible costs related to stress, health effects and service disruption (HCCREMS, 2010) and ecosystems (DCC, 2009; BMT WBM, 2011). Under expected future settlement patterns, exposure of the Australian road and rail network will increase significantly once sea level rises above about 0.5 m (Baynes et al., 2012). Even if temperatures peak and decline, sea level is projected to continue to rise beyond 2100 for many centuries, at a rate dependent on future emissions (AR5 WGI 13.5). Responsibility for adapting to sea level rise in Australasia rests principally with local governments through spatial planning instruments. Western Australia, South Australia and Victoria have mandatory State planning benchmarks for 2100, with local governments determining how they should be implemented. Long-term benchmarks in New South Wales and Queensland have either been suspended or revoked, so local authorities now have broad discretion to develop their own adaptation plans. The New Zealand Coastal Policy Statement (Minister of Conservation, 2010) mandates a minimum 100-year planning horizon for assessing hazard risks, discourages hard protection of existing development and recommends avoidance of new development in vulnerable areas. Non-binding government guidance recommends a risk based approach, using a base value of 0.5 m sea level rise by the 2090s and considering the implications of at least 0.8 m and, for longer term planning, an additional 0.1 m per decade (MfE, 2008d). The incorporation of climate change impacts into local planning has evolved considerably over the past 20 years, but remains piecemeal and shows a diversity of approaches (Gibbs and Hill, 2012; Kay et al., 2014). Governments have invested in high-resolution digital elevation models of coastal and flood prone areas in some regions, but many local governments still lack the resources for hazard mapping and policy design. Political commitment is variable, and legitimacy of approaches and institutions is often strongly contested (Gorddard et al., 2012), including pressure on State governments to modify adaptation policies and on local authorities to compensate developers for restrictions on current or future land uses (LGNZ, 2008; Berry and Vella, 2010; McDonald, 2010; Reisinger et al., 2011). Incremental adaptation responses can entrench existing rights and expectations about on-going protection and development, which limit options for more transformational responses such as accommodation and retreat (high agreement, medium evidence; Gorddard et al., 2012; Barnett et al., 2013; Fletcher et al., 2013; McDonald, 2013). Strategic regional-scale planning initiatives in rapidly growing regions, like south-east Queensland, allow climate change adaptation to be addressed in ways not typically achieved by locality- or sector-specific plans, but require effective coordination across different scales of governance (Serrao-Neumann et al., 2013; Smith et al., 2014). Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Courts in both countries have played an important role in evaluating planning measures. Results of litigation have varied and, in the absence of clearer legislative guidance, more litigation is expected as rising sea levels affect existing properties and adaptation responses constrain development on coastal land (MfE, 2008d; Kenderdine, 2010; Rive and Weeks, 2011; Verschuuren and McDonald, 2012; Corkhill, 2013; Macintosh, 2013). In addition to raising minimum floor levels and creating coastal set-backs to limit further development in areas at risk, several councils in Australia and New Zealand have consulted on or attempted to implement managed retreat policies (ECAN, 2005; BSC, 2010; HDC, 2012; KCDC, 2012). These policies remain largely untested in New Zealand, but experience in Australia has shown high litigation potential and opposing priorities at different levels of government, undermining retreat policies (SCCCWEA, 2009; DCCEE, 2010; Abel et al., 2011). Mandatory disclosure of information about future risks, community engagement and policy stability are critical to support retreat, but existing-use rights, liability concerns, special interests, community resources, place attachment and divergent priorities at different levels of government present powerful constraints (high confidence; Hayward, 2008b; Berry and Vella, 2010; McDonald, 2010; Abel et al., 2011; Alexander et al., 2012; Leitch and Robinson, 2012; Macintosh et al., 2013; Reisinger et al., 2014). _____ END BOX 25-1 HERE _____ 25.4.3. Psychological and Socio-cultural Factors Influencing Impacts of and Adaptation to Climate Change Adapting to climate change relies on individuals accepting and understanding changing risks and opportunities, and responding to these changes both psychologically and behaviourally (see Chapters 2, 16). The majority of Australasians accept the reality of climate change and less than 10% fundamentally deny its existence (high confidence; ShapeNZ, 2009; Leviston et al., 2011; Lewandowsky, 2011; Milfont, 2012; Reser et al., 2012b). Australians perceive themselves to be at higher risk from climate change than New Zealanders and citizens of many other countries, which may reflect recent experiences of climatic extremes (Gifford et al., 2009; Agho et al., 2010; Ashworth et al., 2011; Milfont et al., 2012; Reser et al., 2012c). However, beliefs about climate change and its risks vary over time, are uneven across society and reflect media coverage and bias, political preferences and gender (ShapeNZ, 2009; Bacon, 2011; Leviston et al., 2012; Milfont, 2012), which can influence attitudes to adaptation (Gardner et al., 2010; Gifford, 2011; Reser et al., 2011; Alexander et al., 2012; Raymond and Spoehr, 2013). Surveys in Australia between 2007 and 2011 show moderate to high levels of climate change concern, distress, frustration, resolve, psychological adaptation, and carbon-reducing behaviour (high agreement, medium evidence; Agho et al., 2010; Reser et al., 2012b, c). About two thirds of respondents expected global warming to worsen, with about half very or extremely concerned that they or their family would be affected directly. Direct experience with environmental changes or events attributed to climate change, reported by 45% of respondents, was particularly influential, but the extent to which resulting distress and concern translate into support for planned adaptation has not been fully assessed (Reser et al., 2012a, b). Perceived risks and potential losses from climate change depend on values associated by individuals with specific places, activities and objects. Examples from Australia include the value placed on snow cover in the Snowy Mountains (Gorman-Murray, 2008, 2010), risks to biodiversity and recreational values in coastal South Australia (Raymond and Brown, 2011), conflicts between human uses and environmental priorities in national parks (Wyborn, 2009; Roman et al., 2010), and trade-offs between alternative water supplies and relocation in rural areas (Hurlimann and Dolnicar, 2011). These and additional studies in Australasia confirm that the more individuals identify with particular places and their natural features, the stronger the perceived potential loss but also the greater the motivation to address environmental threats (e.g. Rogan et al., 2005; McCleave et al., 2006; Collins and Kearns, 2010; Gosling and Williams, 2010; Raymond et al., 2011; Russell et al., 2013). This indicates that ecosystem-based climate change adaptation (see Box CC-EA) can provide co-benefits for subjective well-being and mental health, especially for disadvantaged and indigenous communities (Berry et al., 2010; see also 25.8.2). At the same time, social and cultural values and norms can constrain adaptation options for communities by limiting the range of acceptable responses and processes (e.g. place attachment, differing values relating to near- versus long- Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 term, private versus public, and economic versus environmental or social costs and benefits, and perceived legitimacy of institutions). Examples of this are particularly prominent in Australasia in the coastal zone (e.g. Hayward, 2008a; King et al., 2010; Gorddard et al., 2012; Hofmeester et al., 2012) and acceptance of water recycling or pricing (e.g. Pearce et al., 2007; Kouvelis et al., 2010; Mankad and Tapsuwan, 2011). Overall, these studies give high confidence that the experience and threat of climate change and extreme climatic events are having appreciable psychological impacts, resulting in psychological and subsequent behavioural adaptations, reflected in high levels of acceptance and realistic concern, motivational resolve, self-reported changes in thinking, feeling and understanding of climate change and its implications, and behavioural engagement (Reser and Swim, 2011; Reser et al., 2012a, b, c). However, adequate strategies and systems to monitor trends in psychological and social impacts, adaptation and vulnerability are lacking, and such perspectives remain poorly integrated with and dominated by bio-physical and economic characterisations of climate change impacts. 25.5. Freshwater Resources 25.5.1. Observed Impacts Climate change impacts on water represent a cross-cutting issue affecting people, agriculture, industries and ecosystems. The challenge of satisfying multiple demands with a limited resource is exacerbated by the high inter- annual and inter-decadal variability of river flows (Chiew and McMahon, 2002; Peel et al., 2004; Verdon et al., 2004; McKerchar et al., 2010) particularly in Australia. Declining river flows since the mid-1970s in far south- western Australia have led to changed water management (see Box 11.2 in Hennessy et al., 2007). The unprecedented decline in river flows during the 1997-2009 Millennium drought in south-eastern Australia resulted in low irrigation water allocations, severe water restrictions in urban centres, suspension of water sharing arrangements and major environmental impacts (Chiew and Prosser, 2011; Leblanc et al., 2012). 25.5.2. Projected Impacts Figure 25-4 shows estimated changes to mean annual runoff across Australia for a 1°C global average warming above current levels (Chiew and Prosser, 2011; Teng et al., 2012). The range of estimates arises mainly from uncertainty in projected precipitation (Table 25-1). Hydrological modelling with CMIP3 future climate projections indicates that freshwater resources in far south-eastern and far south-west Australia will decline (high confidence; by 0-40% and 20-70%, respectively, for 2°C warming) due to the reduction in winter precipitation (Table 25-1) when most of the runoff in southern Australia occurs. The percent change in mean annual precipitation in Australia is generally amplified as a 2 3 times larger percent change in mean annual streamflow (Chiew, 2006; Jones et al., 2006). This can vary, however, with unprecedented declines in flow in far south-eastern Australia in the 1997 2009 drought (Cai and Cowan, 2008; Potter and Chiew, 2011; Chiew et al., 2013). Higher temperatures and associated evaporation, tree re-growth following more frequent bushfires (Kuczera, 1987; Cornish and Vertessy, 2001; Marcar et al., 2006; Lucas et al., 2007), interceptions from farm dams (van Dijk et al., 2006; Lett et al., 2009) and reduced surface-groundwater connectivity in long dry spells (Petrone et al., 2010; Hughes et al., 2012) can further accentuate declines. In the longer-term, water availability will also be affected by changes in vegetation and surface-atmosphere feedbacks in a warmer and higher CO2 environment (Betts et al., 2007; Donohue et al., 2009; McVicar et al., 2010). [INSERT FIGURE 25-4 HERE Figure 25-4: Estimated changes in mean annual runoff for 1°C global average warming above current levels. Maps show changes in annual runoff (percentage change; top row) and runoff depth (millimetres; bottom row), for dry, median and wet (10th to 90th percentile) range of estimates, based on hydrological modelling using 15 CMIP3 climate projections (Chiew et al., 2009; CSIRO, 2009; Petheram et al., 2012; Post et al., 2012). Projections for 2°C global average warming are about twice that shown in the maps (Post et al., 2011). (Figure adapted from Chiew and Prosser, 2011; Teng et al., 2012).] Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In New Zealand, precipitation changes (Table 25-1) are projected to lead to increased runoff in the west and south of the South Island and reduced runoff in the north-east of the South Island, and the east and north of the North Island (medium confidence). Annual flows of eastward flowing rivers with headwaters in the Southern Alps (Clutha, Waimakariri, Rangitata) are projected to increase by 5-10 % (median projection) by 2040 (Bright et al., 2008; Poyck et al., 2011; Zammit and Woods, 2011) in response to higher alpine precipitation. Most of the increases occur in winter and spring, as more precipitation falls as rain and snow melts earlier (Hendrikx et al., 2013). In contrast, the Ashley River, slightly north of this region, is projected to have little change in annual flows, with the increase in winter flows offset by reduced summer flows (Woods et al., 2008). The retreat of glaciers is expected to have only a minor impact on river flows in the first half of the century (Chinn, 2001; Anderson et al., 2008). Climate change will affect groundwater through changes in recharge rates and the relationship between surface waters and aquifers. Dryland diffuse recharge in most of western, central and southern Australia is projected to decrease because of the decline in precipitation, with increases in the north and some parts of the east because of projected increase in extreme rainfall intensity (medium confidence; Crosbie et al., 2010; McCallum et al., 2010; Crosbie et al., 2012). In New Zealand, a single study projects groundwater recharge in the Canterbury Plains to decrease by about 10% by 2040 (Bright et al., 2008). Climate change will also degrade water quality, particularly through increased material washoff following bushfires and floods (Box 25-6, Box 25-8). 25.5.3. Adaptation The 1997 2009 drought in south-eastern Australia and projected declines in future water resources in southern Australia are already stimulating adaptation (Box 25-2). In New Zealand, there is little evidence of water resources adaptation specifically to climate change. Water in New Zealand is not as scarce generally and water policy reform is driven more by pressure to maintain water quality while expanding agricultural activities, with an increasing focus on collaborative management (Memon and Skelton, 2007; Memon et al., 2010; Lennox et al., 2011; Weber et al., 2011) within national guidelines (LWF, 2010; MfE, 2011). Impacts of climate change on water supply, demand and infrastructure have been considered by several New Zealand local authorities and consultancy reports (Jollands et al., 2007; Williams et al., 2008; Kouvelis et al., 2010), but no explicit management changes have yet resulted. _____ START BOX 25-2 HERE _____ Box 25-2. Adaptation through Water Resources Policy and Management in Australia Widespread drought and projections of a drier future in south-eastern and far south-west Australia (Bates et al., 2010; CSIRO, 2010; Potter et al., 2010; Chiew et al., 2011) saw extensive policy and management change in both rural and urban water systems (Hussey and Dovers, 2007; Bates et al., 2008; Melbourne Water, 2010; DSE, 2011; MDBA, 2011; NWC, 2011; Schofield, 2011). These management changes provide examples of adaptations, building on previous policy reforms (Botterill and Dovers, 2013). The broad policy framework is set out in the 2004 National Water Initiative and 2007 Commonwealth Water Act. The establishment of the National Water Commission (2004) and the Murray-Darling Basin Authority (2008) were major institutional reforms. The National Water Initiative explicitly recognises climate change as a constraint on future water allocations. Official assessments (NWC, 2009, 2011) and critiques (Connell, 2007; Grafton and Hussey, 2007; Byron, 2011; Crase, 2011; Pittock and Finlayson, 2011) have discussed progress and shortcomings of the initiative, but assessment of its overall success is made difficult by other factors such as on-going revisions to allocation plans and time lags to observable impacts. Rural water reform in south-eastern Australia, focused on the Murray-Darling Basin, is currently being implemented. The Murray-Darling Basin Plan (MDBA, 2011, 2012) will return 2750 GL/year of consumptive water (about one fifth of current entitlements) to riverine ecosystems and develop flexible and adaptive water sharing mechanisms to cope with current and future climates. In 2012, the Australian Government committed more than A$12 billion nationally to upgrade water infrastructure, improve water use efficiency, and purchase water entitlements for environmental use. The Basin Plan also includes an environmental watering plan to optimise environmental outcomes for the Basin. Water markets are a key policy instrument, allowing water use patterns to adapt to shifting availability and water to move toward higher value uses (NWC, 2010; Kirby et al., 2012). For Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 example, the two-thirds reduction in irrigation water use over 2000 2009 in the Basin resulted in only 20% reduction in gross agricultural returns, mainly because water use shifted to more valuable enterprises (Kirby et al., 2012). Elsewhere, catchment management authorities and state agencies throughout south-eastern Australia develop water management strategies to cope with prolonged droughts and climate change (e.g. DSE, 2011). Nevertheless, if the extreme dry end of future water projections is realised (25.5.2, Figure 25-4), agriculture and ecosystems across south-eastern and south-western Australia would be threatened even with comprehensive adaptation (see 25.6.1, 25.7.1, 25.7.2; Connor et al., 2009; Kirby et al., 2013). Climate change and population growth are the two major factors that influence water planning in Australian capital cities. In Melbourne, for example, planning has centred on securing new supplies that are more resilient to major climate shocks; increasing use of alternative sources like sewage recycling and stormwater for non-potable water; programs to reduce demand; water sensitive urban design; and integrated planning that considers climate change impact on water supply, flood risk and stormwater and wastewater infrastructures (DSE, 2007; Skinner, 2010; DSE, 2011; Rhodes et al., 2012). Melbourne s water augmentation program includes a desalinisation plant with a 150 GL/year capacity (about one third of the current demand), following the lead of Perth where a desalinisation plant was established in 2006 because of declining inflows since the mid-1970s (Rhodes et al., 2012). Melbourne s water conservation strategies include water efficiency and rebate programs for business and industry, water smart gardens, dual flush toilets, grey water systems, rainwater tank rebates, free water-efficient showerheads and voluntary residential use targets. These conservation measures, together with water use restrictions since the early 2000s, have reduced Melbourne s total per capita water use by 40% (Fitzgerald, 2009; Rhodes et al., 2012). Similar programs reduced Brisbane s per capita water use by about 50% (Shearer, 2011), while adoption of water recycling and rainwater harvesting resulted in up to 60% water savings in some parts of Adelaide (Barton and Argue, 2009). The success of urban water reforms in the face of drought and climate change can be variously interpreted. Increasing supply through desalinisation plants and water reuse schemes reduces the risk of future water shortages and helps cities cope with increasing population. Uptake of household-scale adaptation options has been locally significant but their long-term sustainability or reversibility in response to changing drivers and societal attitudes needs further research (Troy, 2008; Brown and Farrelly, 2009; Mankad and Tapsuwan, 2011). Desalinisation plants can be maladaptive due to their energy demand, and the enhancement of mass supply could create a disincentive for reducing demand or increasing resilience through diversifying supply (Barnett and O'Neill, 2010; Taptiklis, 2011). _____ END BOX 25-2 HERE _____ 25.6. Natural Ecosystems 25.6.1. Inland Freshwater and Terrestrial Ecosystems Terrestrial and freshwater ecosystems have suffered high rates of habitat loss and species extinctions since European settlement in both Australia and New Zealand (Kingsford et al., 2009; Bradshaw et al., 2010; McGlone et al., 2010; Lundquist et al., 2011; SoE, 2011); many reserves are small and isolated, and some key ecosystems and species under-represented (Sattler and Taylor, 2008; MfE, 2010a; SoE, 2011). Many freshwater ecosystems are pressured from over-allocation and pollution, especially in southern and eastern coastal regions in Australia (e.g. Ling, 2010). Additional stresses include erosion, changes in nutrients and fire regimes, mining, invasive species, grazing and salinity (Kingsford et al., 2009; McGlone et al., 2010; SoE, 2011). These increase vulnerability to rapid climate change and provide challenges for both autonomous and managed adaptation (Steffen et al., 2009). 25.6.1.1. Observed Impacts In Australian terrestrial systems, some recently observed changes in the distribution, genetics and phenology of individual species, and in the structure and composition of some ecological communities, can be attributed to recent climatic trends (medium to high confidence; see Box 25-3). Uncertainty remains regarding the role of non-climatic drivers, including changes in atmospheric CO2, fire management, grazing and land-use. The 1997-2009 drought had Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 severe impacts in freshwater systems in the eastern States and the Murray Darling Basin (Pittock and Finlayson, 2011) but in many freshwater systems, direct climate impacts are difficult to detect above the strong signal of over- allocation, pollution, sedimentation, exotic invasions and natural climate variability (Jenkins et al., 2011). In New Zealand, few if any impacts on ecosystems have been directly attributed to climate change rather than variability (Box 25-3; McGlone et al., 2010; McGlone and Walker, 2011). Alpine treelines in New Zealand have remained roughly stable for several hundred years (high confidence) despite 0.9°C average warming over the past century (McGlone and Walker, 2011; Harsch et al., 2012). 25.6.1.2. Projected Impacts Existing environmental stresses will interact with, and in many cases be exacerbated by, shifts in mean climatic conditions and associated change in the frequency or intensity of extreme events, especially fire, drought and floods (high confidence; Steffen et al., 2009; Bradstock, 2010; Murphy et al., 2012). Recent drought-related mortality has been observed for amphibians in south-east Australia (Mac Nally et al., 2009), savannah trees in north-east Australia (Fensham et al., 2009; Allen et al., 2010), mediterranean-type eucalypt forest in southwest Western Australia (Matusik et al., 2013), and, eucalypts in sub-alpine regions in Tasmania (Calder and Kirkpatrick, 2008). Mass die- offs of flying foxes and cockatoos have been observed during heatwaves (Welbergen et al., 2008; Saunders et al., 2011). These examples provide high confidence that extreme heat and reduced water availability, either singly or in combination, will be significant drivers of future population losses and will increase the risk of local species extinctions in many areas (e.g. McKechnie and Wolf, 2010; see also Figure 25-5). Species distribution modelling (SDM) consistently indicates future range contractions for Australia s native species even assuming optimistic rates of dispersal, e.g. Western Australian Banksia spp. (Fitzpatrick et al., 2008), koalas (Adams-Hosking et al., 2011), northern macropods (Ritchie and Bolitho, 2008), native rats (Green et al., 2008b), greater gliders (Kearney et al., 2010b), quokkas (Gibson et al., 2010), platypus (Klamt et al., 2011), birds (Garnett et al., 2013; van der Wal et al., 2013), and fish (Bond et al., 2011). In some studies, complete loss of climatically suitable habitat is projected for some species within a few decades, and therefore increased risk of local and, perhaps, global extinction (medium confidence). SDM has limitations (e.g. Elith et al., 2010; McGlone and Walker, 2011) but is being improved through integration with physiological (Kearney et al., 2010b) and demographic models (Keith et al., 2008; Harris et al., 2012), genetic estimates of dispersal capacity (Duckett et al., 2013), and incorporation into broader risk assessments (e.g. Williams et al., 2008; Crossman et al., 2012). In Australia, assessments of ecosystem vulnerability have been based on observed changes, coupled with projections of future climate in relation to known biological thresholds and assumptions about adaptive capacity (e.g. Laurance et al., 2011; Murphy et al., 2012). There is very high confidence that one of the most vulnerable Australian ecosystems is the alpine zone due to loss of snow cover, invasions by exotic species, and changed species interactions (reviewed in Pickering et al., 2008). There is also high confidence in substantial risks to coastal wetlands such as Kakadu National Park subject to saline intrusion (BMT WBM, 2011); tropical savannas subject to changed fire regimes (Laurance et al., 2011); inland freshwater and groundwater systems subject to drought, over- allocation and altered timing of floods (Pittock et al., 2008; Jenkins et al., 2011; Pratchett et al., 2011); peat-forming wetlands along the east coast subject to drying (Keith et al., 2010); and biodiversity-rich regions such as southwest Western Australia (Yates et al., 2010a; Yates et al., 2010b) and tropical and sub-tropical rainforests in Queensland subject to drying and warming (Stork et al., 2007; Shoo et al., 2011; Murphy et al., 2012; Hagger et al., 2013). The very few studies of climate change impacts on biodiversity in New Zealand suggest that on-going impacts of invasive species (Box 25-4) and habitat loss will dominate climate change signals in the short- to medium-term (McGlone et al., 2010), but that climate change has the potential to exacerbate existing stresses (McGlone and Walker, 2011). There is limited evidence but high agreement that the rich biota of the alpine zone is at risk through increasing shrubby growth and loss of herbs, especially if combined with increased establishment of invasive species (McGlone et al., 2010; McGlone and Walker, 2011). Some cold water-adapted freshwater fish and invertebrates are vulnerable to warming (August and Hicks, 2008; Winterbourn et al., 2008; Hitchings, 2009; McGlone and Walker, 2011) and increased spring flooding may increase risks for braided-river bird species (MfE, 2008b). For some restricted native species, suitable habitat may increase with warming (e.g. native frogs; Fouquet et Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 al., 2010) although limited dispersal ability will limit range expansion. Tuatara populations are at risk as warming increases the ratio of males to females (Mitchell et al., 2010), although the lineage has persisted during higher temperatures in the geological past (McGlone and Walker, 2011). 25.6.1.3. Adaptation High levels of endemism in both countries (Lindenmayer, 2007; Lundquist et al., 2011) are associated with narrow geographic ranges and associated climatic vulnerability, although there is greater scope for adaptive dispersal to higher elevations in New Zealand than in Australia. Anticipated rates of climate change, together with fragmentation of remaining habitat and limited migration options in many regions (Steffen et al., 2009; Morrongiello et al., 2011), will limit in situ adaptive capacity and distributional shifts to more climatically suitable areas for many species (high confidence). Significant local and global losses of species, functional diversity, and ecosystem services, and large scale changes in ecological communities, are anticipated (e.g. Dunlop et al., 2012; Gallagher et al., 2012b; Murphy et al., 2012). There is increasing recognition in Australia that rapid climate change has fundamental implications for traditional conservation objectives (e.g. Steffen et al., 2009; Prober and Dunlop, 2011; Dunlop et al., 2012; Murphy et al., 2012). Research on impacts and adaptation in terrestrial and freshwater systems has been guided by the National Adaptation Research Plans (Hughes et al., 2010; Bates et al., 2011) and by research undertaken within the CSIRO Climate Adaptation Flagship. Climate change adaptation plans developed by many levels of government and Natural Resource Management (NRM) bodies, supported by substantial Australian government funding, have identified priorities that include: identification and protection of climatic refugia (Davis et al., 2013; Reside et al., 2013); restoration of riparian zones to reduce stream temperatures (Davies, 2010; Jenkins et al., 2011); construction of levees to protect wetlands from saltwater intrusion (Jenkins et al., 2011); reduction of non-climatic threats such as invasive species to increase ecosystem resilience (Kingsford et al., 2009); ecologically-appropriate fire regimes (Driscoll et al., 2010); restoration of environmental flows in major rivers (Kingsford and Watson, 2011; Pittock and Finlayson, 2011); protecting and restoring habitat connectivity in association with expansion of the protected area network (Dunlop and Brown, 2008; Mackey et al., 2008; Taylor and Philp, 2010; Prowse and Brook, 2011; Maggini et al., 2013); and, active interventionist strategies such as assisted colonisation to reduce probability of species extinctions (Burbidge et al., 2011; McIntyre, 2011) or restore ecosystem services (Lunt et al., 2013). Few specific measures have been implemented and thus their effectiveness cannot yet be assessed. Biodiversity research and management in New Zealand to date has taken little account of climate change-related pressures and continues to focus largely on managing pressures from invasive species and predators, freshwater pollution, exotic diseases, and halting the decline in native vegetation, although a number of specific recommendations have been made to improve ecosystem resilience to future climate threats (McGlone et al., 2010; McGlone and Walker, 2011). Climate change responses in other sectors may have beneficial as well as adverse impacts on biodiversity, but few tools to assess risks from an integrated perspective have been developed (25.9.1, Box 25-10). Assessments of the impacts of climate change on the provision of ecosystem services (such as pollination and erosion control) via impacts on terrestrial and freshwater ecosystems are generally lacking. Similarly, the concept of Ecosystem-based Adaptation, the role of healthy, well-functioning ecosystems in increasing the resilience of human sectors to the impacts of climate change (see Chapters 4 and 5, and Box CC-EA), is relatively unexplored. 25.6.2. Coastal and Ocean Ecosystems Australia s 60,000 km coastline spans tropical waters in the north to cool temperate waters off Tasmania and the sub-Antarctic islands with sovereign rights over ~8.1 million km2, excluding the Australian Antarctic Territory (Richardson and Poloczanska, 2009). New Zealand has ~18,000 km of coastline, spanning subtropical to sub- Antarctic waters, and the world's fifth largest Exclusive Economic Zone at 4.2 million km2 (Gordon et al., 2010). The marine ecosystems of both countries are considered hotspots of global marine biodiversity with many rare, endemic and commercially important species (Hoegh-Guldberg et al., 2007; Blanchette et al., 2009; Gordon et al., 2010; Gillanders et al., 2011; Lundquist et al., 2011). The increasing density of coastal populations (see 25.3) and Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 stressors such as pollution and sedimentation from settlements and agriculture will intensify non-climate stressors in coastal areas (high confidence; e.g. Russell et al., 2009). Coastal habitats provide many ecosystem services including coastal protection (Arkema et al., 2013) and carbon storage, particularly in seagrass, saltmarsh and mangroves, which could become increasingly important for mitigation (e.g. Irving et al., 2011). Coastal ecosystems occupy <1% of the land mass but may account for 39% of Australia s average national annual carbon burial (estimated total: 466 millions tonnes CO2-eq per year; Lawrence et al., 2012). 25.6.2.1. Observed Impacts There is high confidence that climate change is already affecting the oceans around Australia (Pearce and Feng, 2007; Poloczanska et al., 2007; Lough and Hobday, 2011) and warming the Tasman sea in northern New Zealand (Sutton et al., 2005; Lundquist et al., 2011); average climate zones have shifted south by more than 200 km along the northeast and about 100 km along the northwest Australian coasts since 1950 (Lough, 2008). The rate of warming is even faster in southeast Australia, with a poleward advance of the East Australia Current of ~350 km over the past 60 years (Ridgway, 2007). Based on elevated rates of ocean warming, southwest and southeast Australia are recognized as global warming hotspots (Wernberg et al., 2011). It is virtually certain that the increased storage of carbon by the ocean will increase acidification in the future, continuing the observed trends of the past decades in Australia as elsewhere (Howard et al., 2012; see also WGI 3.8, 6.44). Recently observed changes in marine systems around Australia are consistent with warming oceans (high confidence; Box 25-3). Examples include changes in phytoplankton productivity (Thompson et al., 2009; Johnson et al., 2011); species abundance of macroalgae (Johnson et al., 2011); growth rates of abalone (Johnson et al., 2011), southern rock lobster (Pecl et al., 2009; Johnson et al., 2011), coastal fish (Neuheimer et al., 2011) and coral (De'ath et al., 2009); life cycles of southern rock lobster (Pecl et al., 2009) and seabirds (Cullen et al., 2009; Chambers et al., 2011); and, distribution of subtidal seaweeds (Johnson et al., 2011; Wernberg et al., 2011; Smale and Wernberg, 2013), plankton (Mcleod et al., 2012), fish (Figueira et al., 2009; Figueira and Booth, 2010; Last et al., 2011; Madin et al., 2012), sea urchins (Ling et al., 2009) and intertidal invertebrates (Pitt et al., 2010). Habitat-related impacts are more prevalent in northern Australia (Pratchett et al., 2011), while distribution changes are reported more often in southern waters (Madin et al., 2012), particularly south-east Australia, where warming has been greatest. The 2011 marine heat wave in Western Australia caused the first-ever reported bleaching at Ningaloo reef (Abdo et al., 2012; Feng et al., 2013) resulting in coral mortality (Moore et al., 2012; Depczynski et al., 2013) and changes in community structure and composition (Smale and Wernberg, 2013; Wernberg et al., 2013). About 10% of the observed 50% decline in coral cover on the Great Barrier Reef since 1985 has been attributed to bleaching, the remainder to cyclones and predators (De'ath et al., 2012). Changes in distribution and abundance of marine species in New Zealand are primarily linked to ENSO-related variability that dominates in many time series (Clucas, 2011; Lundquist et al., 2011; McGlone and Walker, 2011; Schiel, 2011), although water temperature is also important (e.g. Beentjes and Renwick, 2001). New Zealand fisheries exported over $1.5 billion worth of product in 2012 (SNZ, 2013) and variability in ocean circulation and temperature plays an important role in local fish abundance (e.g. Chiswell and Booth, 2005; Dunn et al., 2009); no climate change impacts have been reported at this stage (Dunn et al., 2009), although this may be due to insufficient monitoring. 25.6.2.2. Projected Impacts Even though evidence of climate impacts on coastal habitats is limited to date, confidence is high that negative impacts will arise with continued climate change (Lovelock et al., 2009; McGlone and Walker, 2011; Traill et al., 2011; Chapter 6). Some coastal habitats such as mangroves are projected to expand further landward, driven by sea- level rise and exacerbated by soil subsidence if rainfall declines (medium confidence; Traill et al., 2011), although this may be at the expense of saltmarsh and constrained in many regions by the built environment (DCC, 2009; Lovelock et al., 2009; Rogers et al., 2012). Estuarine habitats will be affected by changing rainfall or sediment discharges, as well as connectivity to the ocean (high confidence; Gillanders et al., 2011). Loss of coastal habitats Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and declines in iconic species will result in substantial impacts on coastal settlements and infrastructure from direct impacts such as storm surge, and will affect tourism (medium confidence; 25.7.5). Changes in temperature and rainfall, and sea level rise, are expected to lead to secondary effects, including erosion, landslips, and flooding, affecting coastal habitats and their dependent species, e.g. loss of habitat for nesting birds (high confidence; Chambers et al., 2011). Increasing ocean acidification is expected to affect many taxa (medium confidence; see also Box CC-OA, Chapters 6, 30) including corals (Fabricius et al., 2011), coralline algae (Anthony et al., 2008), calcareous plankton (Richardson et al., 2009; Thompson et al., 2009; Hallegraeff, 2010), reef fishes (Munday et al., 2009; Nilsson et al., 2012), bryozoans, and other benthic calcifiers (Fabricius et al., 2011). Deep-sea scleractinian corals are also expected to decline with ocean acidification (Miller et al., 2011). The AR4 identified the Great Barrier Reef (GBR) as highly vulnerable to both warming and acidification (Hennessy et al., 2007). Recent observations of bleaching (GBRMPA, 2009a) and reduced calcification in both the GBR and other reef systems (Cooper et al., 2008; De'ath et al., 2009; Cooper et al., 2012), along with model and experimental studies (Hoegh-Guldberg et al., 2007; Anthony et al., 2008; Veron et al., 2009) confirm this vulnerability (see also Box CC-CR). The combined impacts of warming and acidification associated with atmospheric CO2 concentrations in excess of 450-500 ppm are projected to be associated with increased frequency and severity of coral bleaching, disease incidence and mortality, in turn leading to changes in community composition and structure including increasing dominance by macroalgae (high confidence; Hoegh-Guldberg et al., 2007; Veron et al., 2009). Other stresses, including rising sea levels, increased cyclone intensity, and nutrient-enriched and freshwater runoff, will exacerbate these impacts (high confidence; Hoegh-Guldberg et al., 2007; Veron et al., 2009; GBRMPA, 2011). Thermal thresholds and the ability to recover from bleaching events vary geographically and between species (e.g. Diaz-Pulido et al., 2009) but evidence of the ability of corals to adapt to rising temperatures and acidification is limited and appears insufficient to offset the detrimental effects of warming and acidification (robust evidence, medium agreement; Hoegh-Guldberg, 2012; Howells et al., 2013; Box CC-CR). Under all SRES scenarios and a range of CMIP3 models, pelagic fishes such as sharks, tuna and billfish are projected to move further south on the east and west coasts of Australia (high confidence; Hobday, 2010). These changes depend on sensitivity to water temperature, and may lead to shifts in species-overlap with implications for by-catch management (Hartog et al., 2011). Poleward movements are also projected for coastal fish species in Western Australia (Cheung et al., 2012) and a complex suite of impacts are expected for marine mammals (Schumann et al., 2013). A strengthening East Auckland Current in northern New Zealand is expected to promote establishment of tropical or sub-tropical species that currently occur as vagrants in warm La Nina years (Willis et al., 2007). Such shifts suggest potentially substantial changes in production and profit of both wild fisheries (Norman-Lopez et al., 2011) and aquaculture species such as salmon, mussels and oysters (medium confidence; Hobday et al., 2008; Hobday and Poloczanska, 2010). Ecosystem models also project changes to habitat and fisheries production (low confidence; Fulton, 2011; Watson et al., 2012). 25.6.2.3. Adaptation In Australia, research on marine impacts and adaptation has been guided by the National Adaptation Research Plan for Marine Biodiversity and Resources (Mapstone et al., 2010), programs within the CSIRO Climate Adaptation Flagship and the Great Barrier Reef Marine Park Authority (GBRMPA, 2007). Limits to autonomous adaptation are unknown for almost all species, although limited experiments suggests capacity for response on a scale comparable to projected warming for some species (e.g. coral reef fish; Miller et al., 2012) and not others (e.g. Antarctic krill; Kawaguchi et al., 2013). Planned adaptation options include removal of human barriers to landward migration of species, beach nourishment, management of environmental flows to maintain estuaries (Jenkins et al., 2010), habitat provision (Hobday and Poloczanska, 2010), assisted colonisation of seagrass and species such as turtles (e.g. Fuentes et al., 2009) and burrow modification for nesting seabirds (Chambers et al., 2011). For southern species on the continental shelf, options are more limited because suitable habitat will not be present the next shallow water to the south is Macquarie Island. There is low confidence about the adequacy of autonomous rates of adaptation by species, although recent experiments with coral reef fish suggest that some species may adapt to the projected climate changes (Miller et al., 2012). Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Management actions to increase coral reef resilience include reducing fishing pressure on herbivorous fish, protecting top predators, managing runoff quality, and minimizing other human disturbances, especially through marine protected areas (Hughes et al., 2007; Veron et al., 2009; Wooldridge et al., 2012). Such actions will slow, but not prevent, long-term degradation of reef systems once critical thresholds of ocean temperature and acidity are exceeded (high confidence), and so novel options, including assisted colonisation and shading critical reefs, have been proposed but remain untested at scale (Rau et al., 2012). Seasonal forecasting can also prepare managers for bleaching events (Spillman, 2011). Adaptation by the fishing industry to shifting distributions of target species is considered possible by most stakeholders (e.g. southern rock lobster fishery; Pecl et al., 2009). Assisted colonisation to maintain production in the face of declining recruitment may also be possible for some high value species, and has been trialled for the southern rock lobster (Green et al., 2010a). Options for aquaculture include disease management, alternative site selection, and selective breeding (Battaglene et al., 2008), but implementation is only preliminary. Marine protected area planning is not explicitly considering climate change in either country, but reserve performance will be affected by projected environment shifts and novel combinations of species, habitats and human pressures (Hobday, 2011). _____ START BOX 25-3 HERE _____ Box 25-3. Impacts of a Changing Climate in Natural and Managed Ecosystems Observed changes in species, and in natural and managed ecosystems (25.6.1, 25.6.2, 25.7.2) provide multiple lines of evidence of the impacts of a changing climate1. Examples of observations published since the AR4 are shown in Table 25-3. [INSERT TABLE 25-3 HERE Table 25-3: Examples of detected changes in species, natural and managed ecosystems, consistent with a climate change1 signal, published since the AR4. Confidence in detection of change is based on the length of study, and the type, amount and quality of data in relation to the natural variability in the particular species or system. Confidence in the role of climate as a major driver of the change is based on the extent to which the detected change is consistent with that expected under climate change, and to which other confounding or interacting non-climate factors have been considered and been found insufficient to explain the observed change.] [FOOTNOTE 1: Consistent with the IPCC definition, a change in climate refers to any statistically detectable signal; it does not necessarily imply a human cause. See Glossary, Table 25-1 and 25.2.] _____ END BOX 25-3 HERE _____ 25.7. Major Industries 25.7.1. Production Forestry Australia has about 149 Mha forests, including woodlands. Two Mha are plantations and 9.4 Mha multiple-use native forests, and forestry contributes around $7 billion annually to GDP (ABARES, 2012). New Zealand s plantation estate in production forests comprises about 1.7 Mha (90% Pinus radiata), with recent contractions due to increased profitability of dairying (FOA and MPI, 2012; MfE, 2013). 25.7.1.1. Observed and Projected Impacts Existing climate variability and other confounding factors have so far prevented the detection of climate change impacts on forests. Modelled projections are based on ecophysiological responses of forests to CO2, water and temperatures. In Australia, potential changes in water availability will be most important (very high confidence; e.g. Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 reviews by Battaglia et al., 2009; Medlyn et al., 2011b). Modelling future distributions or growth rates indicate that plantations in south-west Western Australia are most at risk due to declining rainfall, and there is high confidence that plantation growth will be reduced by temperature increases in hotter regions, especially where species are grown at the upper range of their temperature tolerances (Medlyn et al., 2011a). Moderate reductions in rainfall and increased temperature could be offset by fertilisation from increasing CO2 (limited evidence, medium agreement; Simioni et al., 2009). In cool regions where water is not limiting, higher temperatures could benefit production (Battaglia et al., 2009). In New Zealand, temperatures are mostly sub-optimal for growth of P. radiata and water relations are generally less limiting (Kirschbaum and Watt, 2011). Warming is expected to increase P. radiata growth in the cooler south (very high confidence), whereas in the warmer north, temperature increases can reduce productivity, but CO2 fertilisation may offset this (medium confidence; Kirschbaum et al., 2012). Modelling studies are limited by their reliance on key assumptions which are difficult to verify experimentally, e.g. the degree to which photosynthesis remains stimulated under elevated CO2 (Battaglia et al., 2009). Most studies also exclude impacts of pests, diseases, weeds, fire and wind damage that may change adversely with climate. Fire, for instance, poses a significant threat in Australia and is expected to worsen with climate change (see Box 25-6), especially for the commercial forestry plantations in the southern winter-rainfall regions (Williams et al., 2009; Clarke et al., 2011). In New Zealand, changes in biotic factors are particularly important as they already affect plantation productivity. Dothistroma blight, for instance, is a serious pine disease with a temperature optimum that coincides with New Zealand s warmer, but not warmest, pine-growing regions; under climate change, its severity is, therefore, expected to reduce in the warm central North Island but increase in the cooler South Island (high confidence) where it could offset temperature-driven improved plantation growth (Watt et al., 2011a). There is medium evidence and high agreement of similar future southward shifts in the distribution of existing plantation weed, insect pest and disease species in Australia (see review in Medlyn et al., 2011b). 25.7.1.2. Adaptation Depending on the extent of climate changes and plant responses to increasing CO2, the above studies provide limited evidence but high agreement of potential net increased productivity in many areas, but only where soil nutrients are not limiting. Adaptation strategies include changes to species or provenance selection towards trees better adapted to warmer conditions, or adopting different silvicultural options to increase resilience to climatic or biotic stresses, such as pest challenges (White et al., 2009; Booth et al., 2010; Singh et al., 2010; Wilson and Turton, 2011a). The greatest barriers to long-term adaptation planning are incomplete knowledge of plant responses to increased CO2 and uncertainty in regional climate scenarios (medium evidence, high agreement; Medlyn et al., 2011b). The rotation time of plantation forests of about 30 years or more makes proactive adaptation important but also challenging. 25.7.2. Agriculture Australia produces 93% of its domestic food requirements and exports 76% of agricultural production (PMSEIC, 2010a). New Zealand agriculture contributes about 56% of total export value and dairy products 27%; 95% of dairy products are exported (SNZ, 2012b). Agricultural production is sensitive to climate (especially drought; Box 25-5) but also to many non-climate factors such as management, which thus far has limited both detection and attribution of climate-related changes (see Chapters 7 and 18; Webb et al., 2012a; Darbyshire et al., 2013). Because the region is a major exporter providing, for example, over 40% of the world trade in dairy products changes in production conditions in the region have a major influence on world supply (OECD, 2011). This implies that climate change impacts could have consequences for food security not just locally but even globally (Qureshi et al., 2013a). 25.7.2.1. Projected Impacts and Adaptation Livestock Systems Livestock grazing dominates land use by area in the region. At the Australian national level, the net effect of a 3°C temperature increase (from a 1980-99 baseline) is expected to be a 4% reduction in gross value of the beef, sheep and wool sector (McKeon et al., 2008). Dairy output is projected to decline in all regions of Australia other than Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Tasmania under a 1°C increase by 2030 (Hanslow et al., 2013). Projected changes in national pasture production for dairy, sheep and beef pastures in New Zealand range from an average reduction of 4% across climate scenarios for the 2030s (Wratt et al., 2008) to increases of up to 4% for two scenarios in the 2050s (Baisden et al., 2010) when the models included CO2 fertilisation and nitrogen feedbacks. Studies modelling seasonal changes in fodder supply show greater sensitivity in animal production to climate change and elevated CO2 than models using annual average production, with some impacts expected even under modest warming (high confidence) in both New Zealand (Lieffering et al., 2012) and Australia (Moore and Ghahramani, 2013). Across 25 sites in southern Australia (an area that produces 85% of sheep and 40% of beef production by value) modelled profitability declined at most sites by the 2050s because of a shorter growing season due to changes in both rainfall and temperature (Moore and Ghahramani, 2013). In New Zealand, projected changes in seasonal pasture growth drove changes in animal production at four sites representing the main areas of sheep production (Lieffering et al., 2012). In Hawke s Bay, changes in stock number and the timing of grazing were able to maintain farm income for a period in the face of variable forage supply but not in the longer term. In Southland and Waikato, projected increases in early spring pasture growth posed management problems in maintaining pasture quality, yet, if these were met, animal production could be maintained or increased. The temperature-humidity index (THI), an indicator of potential heat stress for animals, increased from 1960-2008 in the Murray Dairy region of Australia and further increases and reductions in milk production are projected (Nidumolu et al., 2011). Shading can substantially reduce, but not avoid, the temperature and humidity effects that produce a high THI (Nidumolu et al., 2011). Rainfall is a key determinant of inter-annual variability in production and profitability of pastures and rangelands (Radcliffe and Baars, 1987; Steffen et al., 2011) yet remains the most uncertain change. In northern Australia, incremental adaptation may be adequate to manage risks of climate change to the grazing industry but an increasing frequency of droughts and reduced summer rainfall will potentially drive the requirement for transformational change (Cobon et al., 2009). Rangelands that are currently water-limited are expected to show greater sensitivity to temperature and rainfall changes than nitrogen-limited ones (Webb et al., 2012b). The water-sparing effect of elevated CO2 (offsetting reduced water availability from reduced rainfall and increased temperatures) is invoked in many impact studies but does not always translate into production benefits (Kamman et al., 2005; Newton et al., 2006; Stokes and Ash, 2007; Wan et al., 2007). The impacts of elevated CO2 on forage production, quality, nutrient cycling and water availability remains the major uncertainty in modelling system responses (McKeon et al., 2009; Finger et al., 2010); recent findings of grazing impacts on plant species composition (Newton et al., 2013) and nitrogen fixation (Watanabe et al., 2013) under elevated CO2 have added to this uncertainty. New Zealand agro- ecosystems are subject to erosion processes strongly driven by climate; greater certainty in projections of rainfall, particularly storm frequency, are needed to better understand climate change impacts on erosion and consequent changes in the ecosystem services provided by soils (Basher et al., 2012). 25.7.2.2. Projected Impacts and Adaptation Cropping Experiments with elevated CO2 at two sites with different temperatures have shown a wide range in the response of current wheat cultivars (Fitzgerald et al., 2010). Modelling suggests there is the potential to increase New Zealand wheat yields under climate change with appropriate choices of cultivars and sowing dates (high confidence; Teixeira et al., 2012). In Australia, the selection of appropriate cultivars and sowing times are projected to result in increased wheat yields in high rainfall areas such as southern Victoria under climate change and maintenance of current yields in some areas expected to be drier (e.g. north-western Victoria; O'Leary et al., 2010). However, if extreme low rainfall scenarios are realized in areas such as South Australia then changes in cultivars and fertilizer applications are not expected to maintain current yields by 2080 (Luo et al., 2009). Under the more severe climate scenarios and without adaptation, Australia could become a net importer of wheat (Howden et al., 2010). One caveat to modelling studies is that an intercomparison of 27 wheat models found large differences between model outputs for already dry and hot Australian sites in response to increasing CO2 and temperature (Asseng et al., 2013; Carter, 2013). Rice production in Australia is largely dependent on irrigation and climate change impacts will strongly depend on water availability and price (Gaydon et al., 2010). Sugarcane is also strongly water dependent (Carr and Knox, Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2011); yields may increase where rainfall is unchanged or increased, but rising temperatures could drive up evapotranspiration and increase water use (medium confidence; Park et al., 2010). Observed trends and modelling for wine-grapes suggest that climate change will lead to earlier budburst, ripening and harvest for most regions and scenarios (high confidence; Grace et al., 2009; Sadras and Petrie, 2011; Webb et al., 2012a). Without adaptation, reduced quality is expected in all Australian regions (high confidence; Webb et al., 2008). Change in cultivar suitability in specific regions is expected (Clothier et al., 2012), with potential for development of cooler or more elevated sites within some regions (Tait, 2008; Hall and Jones, 2009) and/or expansion to new regions, with some growers in Australia already relocating (e.g. to Tasmania; Smart, 2010). Climate change and elevated CO2 impacts on weeds, pests and diseases are highly uncertain (see Box 25.4). Future performance of currently effective plant resistance mechanisms under elevated CO2 and temperature is particularly important (Melloy et al., 2010; Chakraborty et al., 2011) as is the future efficacy of widely used biocontrol, i.e. the introduction or stimulation of natural enemies to control pests (Gerard et al., 2012). Australia is ranked second and New Zealand fourth in the world in the number of biological control agent introductions (Cock et al., 2010). 25.7.2.3. Integrated Adaptation Perspectives Future water demand by the sector is critical for planning (Box 25-2). Irrigated agriculture occupies less than 1% of agricultural land in Australia but accounted for 28% of gross agricultural production value in 2010/11; almost half of this was produced the Murray Darling Basin, which used 68% of all irrigation water (ABS, 2012b; DAFF, 2012). Reduced inflow under dry climate scenarios is predicted to reduce substantially the value of agricultural production in the Basin (high agreement, robust evidence; Garnaut, 2008; Quiggin et al., 2010; Qureshi et al., 2013b), e.g. in one study by 12-44% to 2030 and 49-72% to 2050 (A1F1; Garnaut, 2008).Water availability also constrains agricultural expansion: 17 Mha in northern Australia could support cropping but only 1% has appropriate water availability (Webster et al., 2009). In New Zealand, the irrigated area has risen by 82% since 1999 to over 1 Mha; 76% is on pasture (Rajanayaka et al., 2010). The New Zealand dairy herd doubled between 1980-2009 expanding from high rainfall zones (>2000 mm annual) into drier, irrigation-dependent areas (600-1000 mm annual); this dependence will increase with further expansion (Robertson, 2010), which is being supported by the Government s Irrigation Acceleration Fund. Many adaptation options such as flexible water allocation, irrigation and seasonal forecasting support managing risk in the current climate (Howden et al., 2008; Botterill and Dovers, 2013) and adoption is often high (Hogan et al., 2011a; Kenny, 2011). However, incremental on-farm adaptation has limits (Park et al., 2012) and may hinder transformational change such as diversification of land use or relocation (see Box 25-5) if it encourages persistence where climate change may take current systems beyond their response capacity (Marshall, 2010; Park et al., 2012; Rickards and Howden, 2012). In many cases, transformational change requires a greater level of commitment, access to more resources, and greater integration across all levels of decision-making that encompass both on- and off-farm knowledge, processes and values (Marshall, 2010; Rickards and Howden, 2012). _____ START BOX 25-4 HERE _____ Box 25-4. Biosecurity Biosecurity is a high priority for Australia and New Zealand given the economic importance of biologically-based industries and risks to endemic species and iconic ecosystems. The biology and potential risk from invasive and native pathogenic species will be altered by climate change (high confidence; Roura-Pascual et al., 2011), but impacts may be positive or negative depending on the particular system. [INSERT TABLE 25-4 HERE Table 25-4: Examples of potential consequences of climate change for invasive and pathogenic species relevant to Australia and New Zealand, with consequence categories based on Hellman et al. (2008).] Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ END BOX 25-4 HERE _____ _____ START BOX 25-5 HERE _____ Box 25-5. Climate Change Vulnerability and Adaptation in Rural Areas Rural communities in Australasia have higher proportions of older and unemployed people than urban populations (Mulet-Marquis and Fairweather, 2008). Employment and economic prospects depend heavily on the physical environment and hence are highly exposed to climate (averages, variability and extremes) as well as changing commodity prices. These interact with other economic, social and environmental pressures, such as changing government policies (e.g. on drought, carbon pricing; Productivity Commission, 2009; Nelson et al., 2010) and access to water resources. The vulnerability of rural communities differs within and between countries reflecting differences in financial security, environmental awareness, policy and social support, strategic skills and capacity for diversification (Bi and Parton, 2008; Marshall, 2010; Nelson et al., 2010; Hogan et al., 2011b; Kenny, 2011). Climate change will affect rural industries and communities through impacts on resource availability and distribution, particularly water. Decreased availability and/or increased demand, or price, in response to climate change will increase tensions among agricultural, mining, urban and environmental water users (very high confidence), with implications for governance and participatory adaptation processes to resolve conflicts (see 25.4.2, 25.6.1, 25.7.2, 25.7.3, Box 25-2, Box 25-10). Communities will also be affected through direct impacts on primary production, extraction activities, critical infrastructure, population health and recreational and culturally significant sites (see 25.7, 25.8; Kouvelis et al., 2010; Balston et al., 2012). Altered production and profitability risks and/or land use will translate into complex and interconnected effects on rural communities, particularly income, employment, service provision, and reduced volunteerism (Stehlik et al., 2000; Bevin, 2007; Kerr and Zhang, 2009). The prolonged drought in Australia during the early 2000s, for example, had many interrelated negative social impacts in rural communities, including farm closures, increased poverty, increased off-farm work and, hence, involuntary separation of families, increased social isolation, rising stress and associated health impacts, including suicide (especially of male farmers), accelerated rural depopulation and closure of key services (high agreement, robust evidence; Alston, 2007; Edwards and Gray, 2009; Alston, 2010, 2012; Hanigan et al., 2012; see also Box CC-GC). Positive social change also occurred, however, including increased social capital through interaction with community organisations (Edwards and Gray, 2009). While social and cultural changes have the potential to undermine the adaptive capacity of communities (Smith et al., 2011), robust ongoing engagement between farmers and the local community can contribute to a strong sense of community and enhance potential for resilience (McManus et al., 2012; see also 25.4.3). The economic impact of droughts on rural communities and the entire economy can be substantial. The most recent drought in Australia (2006/7-2008/9), for example, is estimated to have reduced national GDP by about 0.75% (RBA, 2006) and regional GDP in the southern Murray Darling Basin was about 5.7% below forecast in 2007/08, along with the temporary loss of 6000 jobs (Wittwer and Griffith, 2011). Widespread drought in New Zealand during 2007-2009 reduced direct and off-farm output by about NZ$3.6 billion (Butcher, 2009). The 2012-13 drought in New Zealand is estimated to have reduced national GDP by 0.3-0.6% and contributed to a significant rise in global dairy prices, which tempered even greater domestic economic losses (Kamber et al., 2013). Drought frequency and severity are projected to increase in many parts of the region (Table 25-1). The decisions of rural enterprise managers have significant consequences for and beyond rural communities (Pomeroy, 1996; Clark and Tait, 2008). Many current responses are incremental, responding to existing climate variability (Kenny, 2011). Transformational change has occurred where industries and individuals are relocating part of their operations in response to recent and/or expectations of future climate or policy change (Kenny, 2011; see also Box 25-10), e.g. rice (Gaydon et al., 2010), wine-grapes (Park et al., 2012), peanuts (Thorburn et al., 2012) or changing and diversifying land use in situ (e.g. the recent switch from grazing to cropping in South Australia; Howden et al., 2010). Such transformational changes are expected to become more frequent and widespread with a changing climate (high confidence; 25.7.2), with positive or negative implications for the wider communities in origin and destination regions (Kiem and Austin, 2012). Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Although stakeholders within rural communities differ in their vulnerabilities and adaptive capacities, they are bound by similar dependence upon critical infrastructure and resources, economic conditions, government policy direction, and societal expectations (Loechel et al., 2013). Consequently, adaptation to climate change will require an approach that devolves decision-making to the level where the knowledge for effective adaptations resides, using open communication, interaction and joint-planning (Nelson et al., 2008; Kiem and Austin, 2013). _____ END BOX 25-5 HERE _____ _____ START BOX 25-6 HERE _____ Box 25-6. Climate Change and Fire Fire during hot, dry and windy summers in southern Australia can cause loss of life and substantial property damage (Cary et al., 2003; Adams and Attiwill, 2011). The Black Saturday bushfires in Victoria in February 2009, for example, burnt over 3,500 km2, caused 173 deaths, destroyed over 2,000 buildings and caused damages of A$4 billion (Cameron et al., 2009; VBRC, 2010). This fire occurred toward the end of a 13-year drought (CSIRO, 2010) and after an extended period of consecutive days over 30°C (Tolhurst, 2009). Climate change is expected to increase the number of days with very high and extreme fire weather (Table 25-1), with greater changes where fire is weather-constrained (most of southern Australia; many, in particular eastern and northern, parts of New Zealand) than where it is constrained by fuel load and ignitions (tropical savannas in Australia). Fire season length will be extended in many already high-risk areas (high confidence) and so reduce opportunities for controlled burning (Lucas et al., 2007). Higher CO2 may also enhance fuel loads by increasing vegetation productivity in some regions (Donohue et al., 2009; Williams et al., 2009; Bradstock, 2010; Hovenden and Williams, 2010; King et al., 2011). Climate change and fire will have complex impacts on vegetation communities and biodiversity (Williams et al., 2009). Greatest impacts in Australia are expected in sclerophyll forests of the south-east and south-west (Williams et al., 2009). Most New Zealand native ecosystems have limited exposure but also limited adaptations to fire (Ogden et al., 1998; McGlone and Walker, 2011). There is high confidence that increased fire incidence will increase risk in southern Australia to people, property and infrastructure such as electricity transmission lines (Parsons Brinkerhoff, 2009; O'Neill and Handmer, 2012; Whittaker et al., 2013) and in parts of New Zealand where urban margins expand into rural areas (Jakes et al., 2010; Jakes and Langer, 2012); exacerbate some respiratory conditions such as asthma (Johnston et al., 2002; Beggs and Bennett, 2011); and increase economic risks to plantation forestry (Watt et al., 2008; Pearce et al., 2011). Forest regeneration following wildfires also reduces water yields (Brown et al., 2005; MDBC, 2007), while reduced vegetation cover increases erosion risk and material washoff to waterways with implications for water quality (Shakesby et al., 2007; Wilkinson et al., 2009; Smith et al., 2011a). In Australia, fire management will become increasingly challenging under climate change, potentially exacerbating conflicting management objectives for biodiversity conservation versus protection of property (high confidence; O'Neill and Handmer, 2012; Whittaker et al., 2013). Current initiatives centre on planning and regulations, building design to reduce flammability, fuel management, early warning systems, and fire detection and suppression (Handmer and Haynes, 2008; Preston et al., 2009; VBRC, 2010; O'Neill and Handmer, 2012). Some Australian authorities are taking climate change into account when rethinking approaches to managing fire to restore ecosystems while protecting human life and properties (Preston et al., 2009; Adams and Attiwill, 2011). Improved understanding of climate drivers of fire risk is assisting fire management agencies, landowners and communities in New Zealand (Pearce et al., 2008; Pearce et al., 2011), although changes in management to date show little evidence of being driven by climate change. _____ END BOX 25-6 HERE _____ Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 25.7.3. Mining Australia is the world s largest exporter of coking coal and iron ore and has the world s largest resources of brown coal, nickel, uranium, lead and zinc (ABS, 2012c). Recent events demonstrated significant vulnerability to climate extremes: the 2011 floods reduced coal exports by 25-54 million tonnes and led to A$5-9bn revenue lost in that year (ABARES, 2011; RBA, 2011), and tropical cyclones regularly disrupted mining operations over the past decade (McBride, 2012; Sharma et al., 2013). Flood impacts were exacerbated by regulatory constraints on mine discharges, highlighting tensions among industry, social and ecological management objectives (QRC, 2011), and by flooding affecting road and rail transport to major shipping ports (QRC, 2011; Sharma et al., 2013). Projected changes in climate extremes imply increasing sector vulnerability without adaptation (high confidence; Hodgkinson et al., 2010a; Hodgkinson et al., 2010b). Stakeholders have conducted initial climate risk assessments (Mills, 2009) and perceive the adaptive capacity of the industry to be high (Hodgkinson et al., 2010a; Loechel et al., 2010; QRC, 2011), but costs and broader benefits are yet to be explored along the value-chain and evaluated for community support. On-going challenges include competition for energy and water, climate change scepticism, dealing with contrasting extremes, avoiding maladaptation, and mining-community relations regarding response options, acceptable mine discharges and post-mining rehabilitation (Loechel et al., 2013; Sharma et al., 2013). 25.7.4. Energy Supply, Demand, and Transmission Energy demand is projected to grow by 0.5-1.3% per annum in Australasia over the next few decades in the absence of major new policies (MED, 2011; Syed, 2012). Australia s predominantly thermal power generation is vulnerable to drought-induced water restrictions, which could require dry-cooling and increased water use efficiency where rainfall declines (Graham et al., 2008; Smart and Aspinall, 2009). Depending on carbon price and technology costs, renewable electricity generation in Australia is projected to increase from 10% in 2010/11 to ~33-50% by 2030 (Hayward et al., 2011; Stark et al., 2012; Syed, 2012), but few studies have explored the vulnerability of these new energy sources to climate change (Bryan et al., 2010; Crook et al., 2011; Odeh et al., 2011). New Zealand s predominantly hydroelectric power generation is vulnerable to precipitation variability. Increasing winter precipitation and snow melt, and a shift from snowfall to rainfall will reduce this vulnerability (medium confidence) as winter/spring inflows to main hydro lakes are projected to increase by 5-10% over the next few decades (McKerchar and Mullan, 2004; Poyck et al., 2011). Further reductions in seasonal snow and glacial melt as glaciers diminish, however, would compromise this benefit (Chinn, 2001; Renwick et al., 2009; Srinivasan et al., 2011). Increasing wind power generation (MED, 2011) would benefit from projected increases in mean westerly winds but face increased risk of damages and shut-down during extreme winds (Renwick et al., 2009). Climate warming would reduce annual average peak electricity demands by 1-2% per °C across New Zealand and 2(+/-1)% in New South Wales, but increase by 1.1(+/-1.4)% and 4.6(+/-2.7)% in Queensland and South Australia due to air conditioning demand (Stroombergen et al., 2006; Jollands et al., 2007; Thatcher, 2007; Nguyen et al., 2010). Increased summer peak demand, particularly in Australia (see also Figure 25-5), will place additional stress on networks and can result in black-outs (very high confidence; Jollands et al., 2007; Thatcher, 2007; Howden and Crimp, 2008; Wang et al., 2010a). During the 2009 Victorian heat wave demand rose by 24% but electrical losses from transmission lines increased by 53% due to higher peak currents (Nguyen et al., 2010), and successive failures of the overloaded network temporarily left more than 500,000 people without power (QUT, 2010). Various adaptation options to limit increasing urban energy demand exist and some are being implemented (see Box 25-9). There is limited evidence but high agreement that without additional adaptation, distribution networks in most Australian states will be at high risk of failure by 2031-2070 under non-mitigation scenarios due to increased bushfire risk and potential strengthening and southward shift of severe cyclones in tropical regions (Maunsell and CSIRO, 2008; Parsons Brinkerhoff, 2009). Adaptation costs have been estimated at A$2.5 billion to 2015, with more than half to meet increasing demand for air conditioning and the remainder to increase resilience to climate- related hazards; underground cabling would reduce bushfire risk but has large investment costs that are not included above (Parsons Brinkerhoff, 2009). Decentralised ownership of assets constitutes a significant adaptation constraint Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (ATSE, 2008; Parsons Brinkerhoff, 2009). In New Zealand, increasing high winds and temperatures have been identified qualitatively as the most relevant risks to transmission (Jollands et al., 2007; Renwick et al., 2009). 25.7.5. Tourism Tourism contributes 2.6-4% of GDP to the economies of Australia and New Zealand (ABS, 2010a; SNZ, 2011). The net present value of the Great Barrier Reef alone over the next 100 years has been estimated at A$51.4 billion (Oxford Economics, 2009). Most Australasian tourism is exposed to climate variability and change (see 25.2 for projected trends), and some destinations are highly sensitive to extreme events (Hopkins et al., 2012). The 2011 floods and Tropical Cyclone Yasi, for example, cost the Queensland tourism industry about A$590 million, mainly due to cancellations and damage to the Great Barrier Reef (PWC, 2011); and drought in the Murray-Darling Basin caused an estimated A$70 million loss in 2008 due to reduced visitor days (TRA, 2010). 25.7.5.1. Projected Impacts Future impacts on tourism have been modelled for several Australian destinations. The Great Barrier Reef is expected to degrade under all climate change scenarios (25.6.2, 30.5, Box CC-CR), reducing its attractiveness (Marshall and Johnson, 2007; Bohensky et al., 2011; Wilson and Turton, 2011b). Ski tourism is expected to decline in the Australian Alps due to snow cover reducing more rapidly than in New Zealand (Pickering et al., 2010; Hendrikx et al., 2013) and greater perceived attractiveness of New Zealand (Hopkins et al., 2012). Higher temperature extremes in the Northern Territory are projected with high confidence to increase heat stress and incur higher costs for air conditioning (Turton et al., 2009). Sea level rise places pressures on shorelines and long-lived infrastructure but implications for tourist resorts have not been quantified (Buckley, 2008). Economic modelling suggests that the Australian alpine region would be most negatively affected in relative terms due to limited alternative activities (Pham et al., 2010), whereas the competitiveness of some destinations (e.g. Margaret River in Western Australia) could be enhanced by higher temperatures and lower rainfall (Jones et al., 2010; Pham et al., 2010). An analogue-based study suggests that, in New Zealand, warmer and drier conditions mostly benefit but wetter conditions and extreme climate events undermine tourism (Wilson and Becken, 2011). Confidence in outcomes is low, however, due to uncertain future tourist behaviour (Scott et al., 2012; also 25.9.2). 25.7.5.2. Adaptation Both New Zealand and Australia have formalised adaptation strategies for tourism (Becken and Clapcott, 2011; Zeppel and Beaumont, 2011). In Australia, institutions at various levels also promote preparation for extreme events (Tourism Queensland, 2007, 2010; Tourism Victoria, 2010) and strengthening ecosystem resilience to maintain destination attractiveness (GBRMPA, 2009b). Snow-making is already broadly adopted to increase reliability of skiing (Bicknell and McManus, 2006; Hennessy et al., 2008b), but its future effectiveness depends on location. In New Zealand, even though warming will significantly reduce the number of days suitable for snow-making (Hendrikx and Hreinsson, 2012), sufficient snow could be made in all years until the end of the 21st century to maintain current minimum operational skiing conditions. Options for resorts in Australia s Snowy Mountains are far more limited (Hendrikx et al., 2013), where maintaining skiing conditions until at least 2020 would require A$100 million in capital investment into 700 snow guns and 2.5-3.3 GL of water per month (Pickering and Buckley, 2010). Short investment horizons, high substitutability and a high proportion of human capital compared with built assets give high confidence that the adaptive capacity of the tourism industry is high overall, except for destinations where climate change is projected to degrade core natural assets and diversification opportunities are limited (Evans et al., 2011; Morrison and Pickering, 2011). Strategic adaptation decisions are constrained by uncertainties in regional climatic changes (Turton et al., 2010), limited concern (Bicknell and McManus, 2006), lack of leadership and limited coordinated forward planning (Sanders et al., 2008; Turton et al., 2009; Roman et al., 2010; White and Buultjens, 2012). An integrated assessment of tourism vulnerability in Australasia is not yet possible due to limited Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 understanding of future changes in tourism and community preferences (Scott et al., 2012), including the flow-on effects of changing travel behaviour and tourism preferences in other world regions (see 25.9.2). 25.8.1. Human Health 25.8.1.1. Observed Impacts Life expectancy in Australasia is high, but shows substantial ethnic and socio-economic inequalities (Anderson et al., 2006). Mortality increases in hot weather in Australia (high agreement, robust evidence; Bi and Parton, 2008; Vaneckova et al., 2008) with air pollution exacerbating this association. The last four decades have seen a steady increase in the ratio of summer to winter mortality in Australia, indicating a health effect from climatic warming (Bennett et al., 2013). Exceptional heatwave conditions in Australia have been associated with substantial increases in mortality and hospital admissions in several regional towns and capital cities (high confidence; Khalaj et al., 2010; Loughnan et al., 2010; Tong et al., 2010a; Tong et al., 2010b). For example, during the heatwave in January and February 2009 in south-eastern Australia (BoM, 2009), total emergency cases increased by 46% over the three hottest days. Direct heat-related health problems increased 34-fold, 61% of these being in people aged 75 years or older, and there were an estimated 374 excess deaths, a 62% increase in all-cause mortality (Victorian Government, 2009a). Mental health admissions increased across all age groups by 7.3% in metropolitan South Australia during heatwaves (1993-2006; Hansen et al., 2008). Mortality attributed to mental and behavioural disorders increased in the 65 to 74-year age group and in persons with existing mental health problems (Hansen et al., 2008). Experience of extreme events also strongly affects psychological well-being (see 25.4.3). 25.8.1.2. Projected Impacts Projected increases in heatwaves (Figure 25-5) will increase heat-related deaths and hospitalizations, especially among the elderly, compounded by population growth and ageing (high confidence; Bambrick et al., 2008; Gosling et al., 2009; Huang et al., 2012). In the southern states of Australia and parts of New Zealand, this may be partly offset by reduced deaths from cold at least for modest rises in temperature (low confidence; Bambrick et al., 2008; Kinney, 2012). With strong mitigation, climate change is projected to result in 11% fewer temperature-related deaths in both 2050 and 2100 in Australia, but 14% and 100% more deaths in 2050 and 2100, respectively, without mitigation under a hot, dry A1FI scenario (Bambrick et al., 2008; see Chapter 11 for details on temperature-related health trade-offs). Net results were driven almost entirely by increased mortality in the north, especially Queensland, consistent with Huang et al. (2012). In a separate study that accounted for increased daily temperature variability, a threefold increase in heat-related deaths is projected for Sydney by 2100 for the A2 scenario, assuming no adaptation (Gosling et al., 2009). The number of hot days when physical labor in the sun becomes dangerous is also projected to increase substantially in Australia by 2070, leading to economic costs from lost productivity, increased hospitalisations and occasional deaths (medium confidence; Hanna et al., 2011; Maloney and Forbes, 2011). Water- and food-borne diseases are projected to increase, but the complexity of their relationship to climate and non-climate drivers means there is low confidence in specific projections. For Australia, 205,000-335,000 new cases of bacterial gastroenteritis by 2050, and 239,000-870,000 cases by 2100, are projected under a range of emission scenarios (Bambrick et al., 2008; Harley et al., 2011). Based on their observed positive relationship with temperature, notifications of salmonellosis notifications are projected to increase 15% for every 1°C increase in average monthly temperatures (Britton et al., 2010a). Water-borne zoonotic diseases such as cryptosporidiosis and giardiasis have more complex relationships with climate and are amenable to various adaptations, making future projections more difficult (Britton et al., 2010b; Lal et al., 2012). Understanding of the combined effects of climate change and socio-economic development on the distribution of vector-borne diseases has improved since the AR4. Australasia is projected to remain malaria free under the A1B emission scenario until at least 2050 (Béguin et al., 2011) and sporadic cases could be treated effectively. The area climatically suitable for transmission of dengue will expand in Australasia (high confidence; Bambrick et al., 2008; Aström et al., 2012), but changes in socio-economic factors, especially domestic water-storage may have a more Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 important influence on disease incidence than climate (Beebe et al., 2009; Kearney et al., 2009). Impacts of climate change on Barmah Forest Virus in Queensland depend on complex interactions between rainfall and temperature changes, together with tidal and socio-economic factors, and thus will vary substantially among different coastal regions (Naish et al., 2013). The effects of climate change combined with frequent travel within and outside the region, and recent incursions of exotic mosquito species, could expand the geographic range of other important arboviruses such as Ross River Virus (medium confidence; Derraik and Slaney, 2007; Derraik et al., 2010). A growing literature since the AR4 has focused on the psychological impacts of climate change, based on impacts of recent climate variability and extremes (Doherty and Clayton, 2011; 25.4.3). These studies indicate significant mental health risks associated with climate-related disasters, in particular persistent and severe drought, floods and storms; climate impacts may be especially acute in rural communities where climate change places additional stresses on livelihoods (high confidence; Edwards et al., 2011; see also Box 25-5). Projected population growth and urbanization could further increase health risks indirectly via climate-related stress on housing, transport and energy infrastructure and water supplies (low confidence; Howden-Chapman, 2010; see also Box 25-9). [INSERT FIGURE 25-5 HERE Figure 25-5: Projected changes in exposure to heat under a high emissions scenario (A1FI). Maps show the average number of days with peak temperatures >40°C, for ~1990 (based on available meteorological station data for the period 1975-2004), ~2050 and ~2100. Bar charts show the change in population heat exposure, expressed as person- days exposed to peak temperatures >40°C, aggregated by State/Territory and including projected population growth for a default scenario. Future temperatures are based on simulations by the GFDL-CM2 global climate model (Meehl et al., 2007), re-scaled to the A1FI scenario; simulations based on other climate models could give higher or lower results. Data from Baynes et al. (2012).] 25.8.1.3. Adaptation Research since the AR4 has mainly focused on climate change impacts, although some adaptation strategies have received attention in Australia. These include improving healthcare services, social support for those most at risk, improving community awareness to reduce adverse exposures, developing early warning and emergency response plans (Wang and McAllister, 2011), and understanding perceptions of climatic risks to health as they affect adaptive behaviours (Akompad et al., 2013). In New Zealand, central Government health policies do not identify specific measures to adapt to climate change (Wilson, 2011). In both countries, policies to reduce risks from extreme events such as floods and fires will have co-benefits for health (see Box 25-6, Box 25-8). A review of the southern Australian heatwave of 2009 identified a range of issues including communication failures with no clear public information or warning strategy, and no clear thresholds for initiating public information campaigns (Kiem et al., 2010). Emergency services were underprepared and relied on reactive solutions (QUT, 2010). The Victorian government has since developed a heatwave plan to coordinate a state-wide response, maintain consistent community-wide understanding through a Heat Health alert system, build capacity of councils to support communities most at risk, support a Heat Health Intelligence surveillance system, and distribute public health information (Victorian Government, 2009b). 25.8.2. Indigenous Peoples 25.8.2.1. Aboriginal and Torres Strait Islanders Work since the AR4 includes a national Indigenous adaptation research action plan (Langton et al., 2012), regional risk studies (Green et al., 2009; DNP, 2010; TSRA, 2010; Nursey-Bray et al., 2013) and scrutiny from an Indigenous rights perspective (ATSISJC 2009). Socio-economic disadvantage and poor health (SCRGSP 2011) indicate a disproportionate climate change vulnerability of Indigenous Australians (McMichael et al., 2009) although there are no detailed assessments. In urban and regional areas, where 75% of the Indigenous population lives, assessments have not specifically addressed risks to Indigenous people (e.g. Guillaume et al., 2010). In other Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 regions, all remote, there is limited empirical evidence of vulnerability (Maru et al., 2012). However, there is high agreement and medium evidence for significant future impacts from increasing heat stress, extreme events and increased disease (Campbell et al., 2008; Spickett et al., 2008; Green et al., 2009). The Indigenous estate comprises more than 25% of the Australian land area (Altman et al., 2007; NNTT, 2013). There is high agreement but limited evidence that natural resource dependence (e.g. Bird et al., 2005; Gray et al., 2005a; Kwan et al., 2006; Buultjens et al., 2010) increases Indigenous exposure and sensitivity to climate change (Green et al., 2009); climate change-induced dislocation, attenuation of cultural attachment to place and loss of agency will disadvantage Indigenous mental health and community identity (Fritze et al., 2008; Hunter, 2009; McIntyre-Tamwoy and Buhrich, 2011); and, housing, infrastructure, services and transport, often already inadequate for Indigenous needs especially in remote Australia (ABS 2010c), will be further stressed (Taylor and Philp, 2010). Torres Strait island communities and livelihoods are vulnerable to major impacts from even small sea level rises (high confidence; DCC, 2009; Green et al., 2010b; TSRA 2010). Little adaptation of Indigenous communities to climate change is apparent to date (but see Burroughs, 2010; GETF 2011; Nursey-Bray et al., 2013; Zander et al., 2013). Plans and policies that are imposed on Indigenous communities can constrain their adaptive capacity (Ellemor, 2005; Petheram et al., 2010; Veland et al., 2010; Langton et al., 2012) but participatory development of adaptation strategies is challenged by multiple stressors and uncertainty about causes of observed changes (Leonard et al., 2013b; Nursey-Bray et al., 2013). Adaptation planning would benefit from a robust typology (Maru et al., 2011) across the diversity of Indigenous life experience (McMichael et al., 2009). Indigenous re-engagement with environmental management (e. g. Hunt et al., 2009; Ross et al., 2009) can promote health (Burgess et al., 2009) and may increase adaptive capacity (Berry et al., 2010; Davies et al., 2011). There is emerging interest in integrating Indigenous observations of climate change (Green et al., 2010c; Petheram et al., 2010) and developing inter-cultural communication tools (Woodward et al., 2012; Leonard et al., 2013b). Extensive land ownership in northern and inland Australia and land management traditions mean that Indigenous people are well situated to provide greenhouse gas abatement and carbon sequestration services that may also support their livelihood aspirations (Whitehead et al., 2009; Heckbert et al., 2012). 25.8.2.2. New Zealand M ori The projected impacts of climate change on M ori society are expected to be highly differentiated, reflecting complex economic, social, cultural, environmental and political factors (high confidence). Since the AR4, studies have been either sector-specific (e.g. Insley, 2007; Insley and Meade, 2008; Harmsworth et al., 2010; King et al., 2012) or more general, inferring risk and vulnerability based on exploratory engagements with varied stakeholders and existing social, economic, political and ecological conditions (e.g. MfE, 2007b; Te Aho, 2007; King et al., 2010). The M ori economy depends on climate-sensitive primary industries with vulnerabilities to climate conditions (high confidence; Packman et al., 2001; NZIER, 2003; Cottrell et al., 2004; TPK, 2007; Tait et al., 2008b; Harmsworth et al., 2010; King et al., 2010; Nana et al., 2011a). Much of M ori-owned land is steep (>60%) and susceptible to damage from high intensity rainstorms, while many lowland areas are vulnerable to flooding and sedimentation (Harmsworth and Raynor, 2005; King et al., 2010). Land in the east and north is also drought prone, and this increases uncertainties for future agricultural performance, product quality and investment (medium confidence; Cottrell et al., 2004; Harmsworth et al., 2010; King et al., 2010). The fisheries and aquaculture sector faces substantial risks (and uncertainties) from changes in ocean temperature and chemistry, potential changes in species composition, condition and productivity levels (medium confidence; King et al., 2010; see also 25.6.2). At the community and individual level, M ori regularly utilize the natural environment for hunting and fishing, recreation, the maintenance of traditional skills and identity, and collection of cultural resources (King and Penny, 2006; King et al., 2012). Many of these activities are already compromised due to resource-competition, degradation and modification (Woodward et al., 2001; King et al., 2012). Climate change driven shifts in natural ecosystems will further challenge the capacities of some M ori to cope and adapt (medium confidence; King et al., 2012). M ori organizations have sophisticated business structures, governance (e.g. trusts, incorporations) and networks (e.g. Iwi leadership groups) across the state and private sectors (Harmsworth et al., 2010; Insley, 2010; Nana et al., Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2011b), critical for managing and adapting to climate change risks (Harmsworth et al., 2010; King et al., 2012). Future opportunities will depend on partnerships in business, science, research and government (high confidence; Harmsworth et al., 2010; King et al., 2010) as well as innovative technologies and new land management practices to better suit future climates and use opportunities from climate policy, especially in forestry (Carswell et al., 2002; Harmsworth, 2003; Funk and Kerr, 2007; Insley and Meade, 2008; Tait et al., 2008b; Penny and King, 2010). M ori knowledge of environmental processes and hazards (King et al., 2005; King et al., 2007) as well as strong social-cultural networks are vital for adaptation and on-going risk management (King et al., 2008); however, choices and actions continue to be constrained by insufficient resourcing, shortages in social capital, and competing values (King et al., 2012). Combining traditional ways and knowledge with new and untried policies and strategies will be key to the long-term sustainability of climate-sensitive M ori communities, groups and activities (high confidence; Harmsworth et al., 2010; King et al., 2012). _____ START BOX 25-7 HERE _____ Box 25-7. Insurance as a Climate Risk Management Tool Insurance helps spread the risk from extreme events across communities and over time and therefore enhances the resilience of society to disasters (see 10.7). In Australia, insured losses are dominated by meteorological hazards, including the 2011 Queensland floods and the 1999 Sydney hailstorm (ICA, 2012) with estimated claims of A$3 billion p.a. (IAA, 2011b). In New Zealand, floods and storms are the second most costly natural hazards after earthquakes (ICNZ, 2013). The number of damaging insured events (up to a certain loss value) has increased significantly in the Oceania region since 1980 (Schuster, 2013). Normalised insured losses in Australia show no significant trend from 1967 to 2006 (Crompton and McAneney, 2008; Crompton et al., 2010; Table 10.4), though this conclusion rests on a simplified accounting of population growth and may also reflect improved building codes and early warning systems (Nicholls, 2011; IPCC, 2012). There is high confidence that without adaptive measures, projected increases in extremes (Table 25-1) and uncertainties in these projections will lead to increased insurance premiums, exclusions and non-coverage in some locations (IAG, 2011), which will reshape the distribution of vulnerability, e.g., through unaffordability or unavailability of cover in areas at highest risk (IAA, 2011b, a; NDIR, 2011; Booth and Williams, 2012). Restriction of cover occurred in some locations following the 2011 flood events in Queensland (Suncorp, 2013). Insurance can contribute positively to risk reduction by providing incentives to policy holders to reduce their risk profile (O'Neill and Handmer, 2012), e.g. through resilience ratings given to buildings (TGA, 2009; Edge Environment, 2011; IAG, 2011). Apart from constituting an autonomous private sector response to extreme events, insurance can also be framed as a form of social policy to manage climate risks, similar to New Zealand s government insurance scheme (Glavovic et al., 2010); government measures to reduce or avoid risks also interact with insurance companies willingness to provide cover (Booth and Williams, 2012). Yet insurance can also act as a constraint on adaptation, if those living in climate-risk prone localities pay discounted or cross-subsidised premiums or policies fail to encourage betterment after damaging events by requiring replacement of like for like , constituting a missed opportunity for risk reduction (NDIR, 2011; QFCI, 2012; Reisinger et al., 2014; see also 10.7). The effectiveness of insurance thus depends on the extent to which it is linked to a broader national resilience approach to disaster mitigation and response (Mortimer et al., 2011). _____ END BOX 25-7 HERE _____ _____ START BOX 25-8 HERE _____ Box 25-8. Changes in Flood Risk and Management Responses Flood damages across eastern Australia and both main islands of New Zealand in 2010 and 2011 revealed a significant adaptation deficit (ICA, 2012; ICNZ, 2013). For example, the Queensland floods in January 2011 resulted in 35 deaths, three quarters of the State including Brisbane declared a disaster zone, and damages to public Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 infrastructure of AUD$5-6 billion (Queensland Government, 2011). These floods were associated with a strong monsoon and the strongest La Nina on record (Cai et al., 2012; CSIRO and BoM, 2012; Evans and Boyer-Souchet, 2012). Flood frequency and severity exhibit strong decadal variability with no significant long-term trend in Australasia to date (Kiem et al., 2003; Smart and McKerchar, 2010; Ishak et al., 2013). Flood risk is projected to increase in many regions due to more intense extreme rainfall events driven by a warmer and wetter atmosphere (medium confidence; Table 25-1). High resolution downscaling (Carey-Smith et al., 2010), and dynamic catchment hydrological and river hydraulic modelling in New Zealand (Gray et al., 2005b; McMillan et al., 2010; MfE, 2010b; Ballinger et al., 2011; Duncan and Smart, 2011; McMillan et al., 2012) indicate that the 50-year and 100-year flood peaks for rivers in many parts of the country will increase by 5 10% by 2050 and more by 2100 (with large variation between models and emissions scenarios), with a corresponding decrease in return periods for specific flood levels. Studies for Queensland show similar results (DERM et al., 2010). In Australia, flood risk is expected to increase more in the north (driven by convective rainfall systems) than in the south (where more intense extreme rainfall may be compensated by drier antecedent moisture conditions), consistent with confidence in heavy rainfall projections (Table 25-1; Alexander and Arblaster, 2009; Rafter and Abbs, 2009). Flood risk near river mouths will be exacerbated by storm surge associated with higher sea level and potential change in wind speeds (McInnes et al., 2005; MfE, 2010b; Wang et al., 2010b). Higher rainfall intensity and peak flow will also increase erosion and sediment loads in waterways (Prosser et al., 2001; Nearing et al., 2004) and exacerbate problems from aging stormwater and wastewater infrastructure in cities (Howe et al., 2005; Jollands et al., 2007; CCC, 2010; WCC, 2010; see also Box 25-9). However, moderate flooding also has benefits through filling reservoirs, recharging groundwater and replenishing natural environments (Hughes, 2003; Chiew and Prosser, 2011; Oliver and Webster, 2011). Adaptation to increased flood risk from climate change is starting to happen (Wilby and Keenan, 2012) through updating guidelines for design flood estimation (MfE, 2010b; Westra, 2012), improving flood risk management (O'Connell and Hargreaves, 2004; NFRAG, 2008; Queensland Government, 2011), accommodating risk in flood prone areas (options include raising floor levels, using strong piled foundations, using water-resistant insulation materials and ensuring weather tightness), and risk reduction and avoidance through spatial planning and managed relocation (Trotman, 2008; Glavovic et al., 2010; LVRC, 2012; QFCI, 2012). Adaptation options in urban areas also include ecosystem-based approaches such as retaining floodplains and floodways, restoring wetlands, and retrofitting existing systems to attenuate flows (Box 25.9; Howe et al., 2005; Skinner, 2010; WCC, 2010). The recent flooding in eastern Australia and the projected increase in future flood risk have resulted in changes to reservoir operations to mitigate floods (van den Honert and McAneney, 2011; QFCI, 2012) and insurance practice to cover flood damages (NDIR, 2011; Phelan, 2011; Box 25-7). However, the magnitude of potential future changes in flood risk and limits to incremental adaptation responses in urban areas suggest that more transformative approaches based on altering land-use and avoidance of exposure to future flooding may be needed in some locations, especially if changes in the upper range of projections are realised (high confidence; Lawrence and Allan, 2009; DERM et al., 2010; Glavovic et al., 2010; Wilby and Keenan, 2012; Lawrence et al., 2013a). _____ END BOX 25-8 HERE _____ 25.9. Interactions among Impacts, Adaptation, and Mitigation Responses The AR4 found that individual adaptation responses can entail synergies or trade-offs with other adaptation responses and with mitigation, but that integrated assessment tools were lacking in Australasia (Hennessy et al., 2007). Subsequent studies provide detail on such interactions and can inform a balanced portfolio of climate change responses, but evaluation tools remain limited, especially for local decision-making (Park et al., 2011). A review of 25 specific climate change-associated land-use plans from Australia, for example, found that 14 exhibited potential for conflict between mitigation and adaptation (Hamin and Gurran, 2009). Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ START BOX 25-9 HERE _____ Box 25-9. Opportunities, Constraints, and Challenges to Adaptation in Urban Areas Considerable opportunities exist for Australasian cities and towns to reduce climate change impacts and, in some regions, benefit from projected changes such as warmer winters and more secure water supply (Fitzharris, 2010; Australian Government, 2012). Many tools and practices developed for sustainable resource management or disaster risk reduction in urban areas are co-beneficial for climate change adaptation, and vice versa, and can be integrated with mitigation objectives (Hamin and Gurran, 2009). Despite the abundance of potential adaptation options, however, social, cultural, institutional and economic factors frequently constrain their implementation (high confidence; see also 25.4.2). The form and longevity of cities and towns, with their concentration of hard and critical infrastructure such as housing, transport, energy, stormwater and wastewater systems, telecommunications and public facilities provide additional challenges (see also Chapters 8 and 10, 25.7.4, 25.8.1, Boxes 25-1, 25-2, 25-8). Transport infrastructure is vulnerable to extreme heat and flooding (QUT, 2010; Taylor and Philp, 2010) but quantification of future risks remains limited (Gardiner et al., 2009; Balston et al., 2012; Baynes et al., 2012). Table 25-5 summarises some adaptation options, co-benefits and constraints on their adoption in Australasia. Overall, the implementation of climate change adaptation policy for urban settlements in Australia and New Zealand has been mixed. The Australian National Urban Policy encourages adaptation, and many urban plans include significant adaptation policies (e.g. City of Melbourne, 2009; City of Port Phillip, 2010; ACT Government, 2012; City of Adelaide, 2012). New Zealand also promotes urban adaptation through strategies, plans and guidance documents (MfE, 2008b; CCC, 2010; WCC, 2010; Auckland Council, 2012; NIWA et al., 2012). Many examples of incremental urban adaptation exist (Box 25-2, Table 25-5), particularly where these include co-benefits and respond to other stressors, like prolonged drought in southern Australia and recurrent floods. Experience is much scarcer with more flexible land-uses, managed relocation and ecosystem-based adaptation that could transform existing settlement patterns and development trends, and where maintaining flexibility to address long-term climate risks can run against near-term development pressures (see Boxes 25-1, 25-2, 25-8, CC-EA). Decision-making models that support such adaptive and transformative changes (25.4.2, Box 25-1) have not yet been implemented widely in urban contexts; increased coordination among different levels of government may be required to spread costs and balance public and private, near- and long-term and local and regional benefits (Norman, 2009; Britton, 2010; Norman, 2010; Abel et al., 2011; Lawrence et al., 2013a; McDonald, 2013; Palutikof et al., 2013; Reisinger et al., 2014). [INSERT TABLE 25-5 HERE Table 25-5: Examples of co-beneficial climate change adaptation options for urban areas and barriers to their adoption. Options in italics are already widely implemented in Australia and New Zealand urban areas.] _____ END BOX 25-9 HERE _____ 25.9.1. Interactions among Local-Level Impacts, Adaptation, and Mitigation Responses Table 25-6 shows examples of adaptation responses that are either synergistic or entail trade-offs with other impacts and/or adaptation responses and goals. Adapting proactively to projected climate changes, particularly extremes such as floods or drought, can increase near-term resilience to climate variability and be a motivation for adopting adaptation measures (Productivity Commission, 2012). However, exclusive reliance on near-term benefits can increase trade-offs and result in long-term maladaptation (high confidence). For example, enhancing protection measures after major flood events, combined with rapid re-building, accumulates fixed assets that can become increasingly costly to protect as climate change continues, with attendant loss of amenity and environmental values (Glavovic et al., 2010; Gorddard et al., 2012; McDonald, 2013). Similarly, deferring adoption of increased design wind speeds in cyclone-prone areas delays near-term investment costs but also reduces the long-term benefit/cost ratio of the strategy (Stewart and Wang, 2011). Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT TABLE 25-6 HERE Table 25-6: Examples of interactions between impacts and adaptation measures in different sectors. In each case, impacts or responses in one sector have the potential to cause negative impacts or have co-benefits with impacts or responses in another sector, or with another type of response in the same sector.] Mitigation actions can contribute to but also counteract local adaptation goals. Energy efficient buildings, for example, reduce network and health risks during heat waves, but urban densification to reduce transport energy demand intensifies urban heat islands and, hence, heat-related health risks (25.7.4, 25.8.1). Specific adaptations can also make achievement of mitigation targets harder or easier. Increased use of air conditioning, for example, increases energy demand, but energy efficiency and building design can reduce heat exposure as well as energy demand (25.7.4, Box 25-9). Table 25-7 gives further examples, and Box 25-10 explores the multiple and complex benefits and trade-offs in changing land-use to simultaneously adapt to and mitigate climate change. [INSERT TABLE 25-7 HERE Table 25-7: Examples of interactions between adaptation and mitigation measures (green rows denote synergies where multiple benefits may be realized, orange rows denote potential tradeoffs and conflicts; grey row gives an example of complex, mixed interactions). The primary goal may be adaptation or mitigation.] _____ START BOX 25-10 HERE _____ Box 25-10. Land-based Interactions Among Climate, Energy, Water, and Biodiversity Climate, water, biodiversity, food and energy production and use are intertwined through complex feedbacks and trade-offs (see also Box CC-WE). This could make alternative uses of natural resources within rural landscapes increasingly contested, yet decision support tools to manage competing objectives are limited (PMSEIC, 2010b). Various policies in Australasia support increased biofuel production and biological carbon sequestration via, for example, mandatory renewable energy targets and incentives to increase carbon storage. Impacts of increased biological sequestration activities on biodiversity depend on their implementation. Benefits arise from reduced erosion, additional habitat, and enhanced ecosystem connectivity, while risks or lost opportunities are associated with large-scale monocultures especially if replacing more diverse landscapes (Brockerhoff et al., 2008; Giltrap et al., 2009; Steffen et al., 2009; Todd et al., 2009; Bradshaw et al., 2013). Photosynthesis transfers water to the atmosphere, so increased sequestration is projected to reduce catchment yields particularly in southern Australia and affect water quality negatively (CSIRO, 2008; Schrobback et al., 2011; Bradshaw et al., 2013). Accounting for this water use in water allocations for sequestration activities would increase their cost and limit the potential of sequestration-driven land-use change (Polglase et al., 2011; Stewart et al., 2011). Large-scale land-cover changes also affect local and regional climates and soil moisture through changing albedo, evaporation, plant transpiration and surface roughness (McAlpine et al., 2009; Kirschbaum et al., 2011b), but these feedbacks have rarely been included in analyses of changing water demands and availability. Biological carbon sequestration in New Zealand is less water-challenged than in Australia, except where catchments are projected to become drier and/or are already completely allocated (MfE, 2007a; Rutledge et al., 2011), and would mostly improve water quality through reduced erosion (Giltrap et al., 2009). Policies to protect water quality by limiting nitrogen discharge from agriculture have reduced livestock production and greenhouse gas emissions in the Lake Taupo and Rotorua catchments and supported land-use change towards sequestration (OECD, 2013b). Trade-offs between biofuel and food production and ecosystem services depend strongly on the type of sequestration activity and their management relies on the use of consistent principles to evaluate externalities and benefits of alternative land-uses (PMSEIC, 2010b). First-generation biofuels have been modelled in Australia as directly competing with agricultural production (Bryan et al., 2010). In contrast, production of woody biofuels in New Zealand is projected to occur on marginal land, not where the most intense agriculture occurs (Todd et al., 2009). Falling costs and increasing efficiency of solar energy may limit future biofuel demand, given the limited efficiency of plants in converting solar energy into usable fuel (e.g. Reijnders and Huijbregts, 2007). Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 _____ END BOX 25-10 HERE _____ 25.9.2. Intra- and Inter-Regional Flow-On Effects Among Impacts, Adaptation and Mitigation Recent studies strengthen conclusions from the AR4 (Hennessy et al., 2007) that flow-on effects from climate change impacts occurring in other world regions can exacerbate or counteract projected impacts in Australasia. Modelling suggests Australia s terms of trade would deteriorate by about 0.23% in 2050 and 2.95% in 2100 as climate change impacts without mitigation reduce economic activity and demand for coal, minerals and agricultural products in other world regions (A1FI scenario; Harman et al., 2008). As a result, Australian Gross National Product (GNP) is expected to decline more strongly than GDP due to climate change, especially towards the end of the 21st century (Gunasekera et al., 2008). These conclusions, however, merit only medium confidence, because they rely on simplified assumptions about global climate change impacts, economic effects and policy responses. For New Zealand, there is limited evidence but high agreement that higher global food prices driven by adverse climate change impacts on global agriculture and some international climate policies would increase commodity prices and hence producer returns. Agriculture and forestry producer returns, for example, are estimated to increase by 14.6% under the A2 scenario by 2070 (Saunders et al., 2010) and real gross national disposable income by 0.6- 2.3% under a range of non-mitigation scenarios (Stroombergen, 2010) relative to baseline projections in the absence of global climate change. Some climate policies such as biofuel targets and agricultural mitigation in other regions would also increase global commodity prices and hence returns to New Zealand farmers (Saunders et al., 2009; Reisinger et al., 2012). Depending on global implementation, these could more than offset projected average domestic climate change impacts on agriculture (Tait et al., 2008a). In contrast, higher international agricultural commodity prices appear insufficient to compensate for the more severe effects of climate change on agriculture in Australia (see 25.7.2; Gunasekera et al., 2007; Garnaut, 2008). Climate change could affect international tourism to Australasia through international destination and activity preferences (Kulendran and Dwyer, 2010; Rosselló-Nadal et al., 2011; Scott et al., 2012), climate policies, and oil prices (Mayor and Tol, 2007; Becken, 2011; Schiff and Becken, 2011). These potentially significant effects remain poorly quantified, however, and are not well integrated into local vulnerability studies (Hopkins et al., 2012). Climate change has the potential to change migration flows within Australasia, particularly due to coastal changes (e.g. from the Torres Straits islands to mainland Australia), although reliable estimates of such movements do not yet exist (see 12.4; Green et al., 2010b; McNamara et al., 2011; Hugo, 2012). Migration within countries, and from New Zealand to Australia, is largely economically driven and sustained by transnational networks, though the perceived more attractive current climate in Australia is reportedly a factor in migration from New Zealand (Goss and Lindquist, 2000; Green et al., 2008a; Poot, 2009). The impacts of climate change in the Pacific may contribute to an increase in the number of people seeking to move to nearby countries (Bedford and Bedford, 2010; Hugo, 2010; McAdam, 2010; Farbotko and Lazrus, 2012; Bedford and Campbell, 2013) and affect political stability and geopolitical rivalry within the Asia-Pacific region, although there is no clear evidence of this to date and causal theories are scarce (see 12.4, 12.5; Dupont, 2008; Pearman, 2009). Increasing climate-driven disasters, disease and border control will stimulate operations other than war for Australasia s armed forces; integration of security into adaptation and development assistance for Pacific island countries can therefore play a key role in moderating the influence of climate change on forced migration and conflict (high agreement, robust evidence; Dupont and Pearman, 2006; Bergin and Townsend, 2007; Dupont, 2008; Sinclair, 2008; Barnett, 2009; Rolfe, 2009). 25.10. Synthesis and Regional Key Risks 25.10.1. Economy-wide Impacts and the Potential of Mitigation to Reduce Risks Globally effective mitigation could reduce or delay some of the risks associated with climate change and make adaptation more feasible beyond about 2050, when projected climates begin to diverge substantially between Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 mitigation and non-mitigation scenarios (see also 19.7). Literature quantifying these benefits for Australasia has increased since the AR4 but remains very sparse. Economy-wide net costs for Australia are modelled to be substantially greater in 2100 under unmitigated climate change (A1FI; GNP loss 7.6%) than under globally effective mitigation (GNP loss less than 2% for stabilization at 450 or 550 ppm CO2-eq, including costs of mitigation and residual impacts; Garnaut, 2008). These estimates, however, are highly uncertain and depend strongly on valuation of non-market impacts, treatment of potentially catastrophic outcomes, and assumptions about adaptation, global changes and flow-on effects for Australia, and effectiveness and implementation of global mitigation efforts (Garnaut, 2008). No estimates of climate change costs across the entire economy exist for New Zealand. The benefits of mitigation in terms of reduced risks have been quantified for some individual sectors in Australia, e.g. for irrigated agriculture in the Murray-Darling Basin (Quiggin et al., 2008; Quiggin et al., 2010; Valenzuela and Anderson, 2011; Scealy et al., 2012) and for net health outcomes (Bambrick et al., 2008). Although quantitative estimates from individual studies are highly assumption-dependent, multiple lines of evidence (see 25.7, 25.8) give very high confidence that globally effective mitigation would significantly reduce many long-term risks from climate change to Australia. Benefits differ, however, between States for some issues, e.g. heat and cold mortality (Bambrick et al., 2008). Few studies consider mitigation benefits explicitly for New Zealand, but scenario-based studies give high confidence that if global emissions were reduced from a high (A2) to a medium-low (B1) emissions scenario, this would markedly lower the projected increase in flood risks (Ballinger et al., 2011; McMillan et al., 2012) and reduce risks to livestock production in the most drought prone regions (Tait et al., 2008a; Clark et al., 2011). Mitigation would also reduce the projected benefits to production forestry, however, though amounts depend on the response to CO2 fertilization (Kirschbaum et al., 2011a; 25.7.1). 25.10.2. Regional Key Risks as a Function of Mitigation and Adaptation The Australia/New Zealand Chapter of the AR4 (Hennessy et al., 2007) concluded with an assessment of aggregated vulnerability for a range of sectors as a function of global average temperature. Building on recent additional insights, Table 25-8 shows eight key risks within those sectors that can be identified with high confidence for the 21st century, based on the multiple lines of evidence presented in the preceding sections and selected using the framework for identifying key risks set out in Chapter 19 (see also Box CC-KR). This combines consideration of biophysical impacts, their likelihood, timing and persistence, with vulnerability of the affected system, based on exposure, magnitude of harm, significance of the system and its ability to cope with or adapt to projected biophysical changes. These key risks differ in the extent to which they can be managed through adaptation and mitigation and their evolution over time, and some are more likely than others, but all warrant attention from a risk- management perspective. [INSERT TABLE 25-8 HERE Table 25-8: Key regional risks from climate change and the potential for reducing risk through mitigation and adaptation. Key risks are identified based on assessment of the literature and expert judgments by chapter authors, with evaluation of evidence and agreement in the supporting chapter sections. Each key risk is characterized on a scale from very low to very high and presented in three timeframes: the present, near-term (2030-2040), and long- term (2080-2100). For the near-term era of committed climate change (here, for 2030-2040), projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. For the longer-term era of climate options (here, for 2080-2100), risk levels are presented for global mean temperature increase of 2°C and 4°C above preindustrial levels. For each timeframe, risk levels are estimated for a continuation of current adaptation and for a hypothetical highly adapted state. Relevant climate variables are indicated by icons. For a given key risk, change in risk level through time and across magnitudes of climate change is illustrated, but because the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks, sectors, or regions.] One set of key risks comprises damages to natural ecosystems (significant change in community structure of coral reefs and loss of some montane ecosystems) that can be moderated by globally effective mitigation but to which some damage now seems inevitable. For some species and ecosystems, climatically constrained ecological niches, fragmented habitats and limited adaptive movement collectively present hard limits to adaptation to further climate Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 change (high confidence). A second set of key risks (increase in flood risk, water scarcity, heat waves and wild fire) comprises damages that could be severe but can be reduced substantially by globally effective mitigation combined with adaptation, with the need for transformational adaptation increasing with the rate and amount of climate change. A third set of key risks (coastal damages from sea level rise, and loss of agriculture production from severe drying) comprises potential impacts whose scale remains highly uncertain within the 21st century, even for a given global temperature change, and where alternative scenarios materially affect levels of concern, adaptation needs and strategies. Even though scenarios of severe drying (see 25.5.2) or rapid sea level rise approaching 1 m or more by 2100 (see Box 25-2 and WGI 13.5) have low or currently unknown probabilities, the associated impacts would so severely challenge adaptive capacity, including transformational changes, that they constitute important risks. A first comparative assessment for Australia of exposure and damages from different hazards up to 2100 indicates that river flooding will continue to be the most costly source of direct damages to infrastructure, even though the largest value of assets is exposed to bush fire. Exposure to and damages from coastal inundation are currently smaller, but would rise most rapidly beyond mid-century if sea level rise exceeds 0.5 m (Baynes et al., 2012). An emerging risk is the compounding of extreme events, none of which would constitute a key risk in its own right, but that collectively and cumulatively across space and time could stretch emergency response and recovery capacity and hamper regional economic development, including through impacts on insurance markets or multiple concurrent needs for major infrastructure upgrades (NDIR, 2011; Phelan, 2011; Baynes et al., 2012; Booth and Williams, 2012; Karoly and Boulter, 2013). Efforts are underway to better understand the potential importance of cumulative impacts and responses, including the challenges arising from impacts and responses across different levels of government (CSIRO, 2011; Leonard et al., 2013a), but evidence is as yet too limited to identify this as a key risk consistent with the definitions adopted in this report (see Chapter 19). Climate change is projected to bring benefits to some sectors and parts of Australasia, at least under limited warming scenarios associated with globally effective mitigation (high confidence). Examples include an extended growing season for agriculture and forestry in cooler parts of New Zealand and Tasmania, reduced winter mortality (low confidence) and reduced winter energy demand in most of New Zealand and southern States of Australia, and increased winter hydropower potential in New Zealand s South Island (25.7.1, 25.7.2, 25.7.4, 25.8.1). The literature supporting this assessment of key risks is uneven among sectors and between Australia and New Zealand; for the latter, conclusions in many sectors are based on limited studies that often use a narrow set of assumptions, models, and data and which, accordingly, have not explored the full range of potential outcomes. 25.10.3. Challenges to Adaptation in Managing Key Risks, and Limits to Adaptation Two key and related challenges for regional adaptation are apparent: to identify when and where adaptation may imply transformational rather than incremental changes; and, where specific interventions are needed to overcome adaptation constraints, in particular to support transformational responses that require coordination across different spheres of governance and decision-making (Productivity Commission, 2012; Palutikof et al., 2013). The magnitude of climate change, especially under scenarios of limited mitigation, and constraints to adaptation suggest that incremental and autonomous responses will not deliver the full range of available adaptation options nor ensure the continued function of natural and human systems if some key risks are realized (high confidence; see also 25.4). Most incremental adaptation measures in natural ecosystems focus on reducing other non-climate stresses but, even with scaled-up efforts, conserving the current state and composition of the ecosystems most at risk appears increasingly infeasible (25.6.1, 25.6.2). Maintenance of key ecosystem functions and services requires a radical reassessment of conservation values and practices related to assisted colonisation and the values placed on introduced species (Steffen et al., 2009). Divergent views regarding intrinsic and service values of species and ecosystems imply the need for a proactive discussion to enable effective decision-making and resource allocation. In human systems, incremental adjustments of current risk management tools, planning approaches and early warning systems for floods, fire, drought, water resources and coastal hazards can increase resilience to climate Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 variability and change especially in the near term (IPCC, 2012; Productivity Commission, 2012; Dovers, 2013). A purely incremental approach, however, which generally aims to preserve current management objectives, governance and institutional arrangements, can make later transformational changes increasingly difficult and costly (high agreement, medium evidence; e.g. Howden et al., 2010; Park et al., 2012; McDonald, 2013; Stafford-Smith, 2013). Examples of transformational changes include: shifting emphasis from protection to accommodation or avoidance of flood risk, including managed retreat from eroding coasts; the translocation of industries in response to increasing drought, flood and fire risks or water scarcity; and the associated transformation of the economic and social base and governance of some rural communities (Boxes 25-1, -2, 5-9; Nelson et al., 2010; Linnenluecke et al., 2011; Kiem and Austin, 2012; O'Neill and Handmer, 2012; McDonald, 2013; Palutikof et al., 2013). Consideration of transformational adaptation becomes critical where long life- or lead-times are involved, and where high up-front costs or multiple interdependent actors create constraints that require coordinated and proactive interventions (Stafford-Smith et al., 2011; Productivity Commission, 2012; Palutikof et al., 2013). Deferring such adaptation decisions due to uncertainty about the future will not necessarily minimize costs or ensure adequate flexibility for future responses, although up-front investment and opportunity costs of adaptation can present powerful arguments for delayed or staged responses (Stewart and Wang, 2011; Gorddard et al., 2012; Productivity Commission, 2012; McDonald, 2013). Whether transformational responses are seen as success or failure of adaptation depends on the extent to which actors accept a change in, or wish to maintain current activities and management objectives, and the degree to which the values and institutions underpinning the transformation are shared or contested across stakeholders (Park et al., 2012; Stafford-Smith, 2013). These views will differ not only between communities and industries but also from person to person depending on their individual value systems, perceptions of and attitude to risk, and ability to capitalize on opportunities (see also 25.4.3). 25.11. Filling Knowledge Gaps to Improve Management of Climate Risks The wide range of projected rainfall changes (averages and extremes) and their hydrological amplification are key uncertainties affecting the scale and urgency of adaptation in agriculture, forestry, water resources, some ecosystems, and wildfire and flood risks. For ecosystems, agriculture and forestry, these uncertainties are compounded by limited knowledge of responses of vegetation to elevated CO2, changes in ocean pH, and interactions with changing climatic conditions. The uncertainties in future impacts are most critical for decisions with long lifetimes, such as capital infrastructure investment or large-scale changes in land- and water-use. Uncertainties about the rate of sea level rise, and changes in storm paths and intensity, add to challenges for infrastructure design. The use of multi-model means and a narrow set of emissions scenarios in many past studies implies that the full set of climate-related risks and management options remains incompletely explored. Understanding of ecological and physiological thresholds that, once exceeded, would result in rapid changes in species, ecosystems and their services, is still very limited. The literature is noticeably sparse in New Zealand and for arid Australia. These knowledge gaps are compounded by limited information about the effect of global climate change on patterns of natural climate variability, such as ENSO. Better understanding the effect of evolving natural climate variability and long-term trends, along with rising CO2 concentrations, on pests, invasive species and native and managed ecosystems could support more robust ecosystem-based adaptation strategies. Vulnerability of human and managed systems depends critically on future socio-economic characteristics. Research into psychological, economic, social and cultural dimensions of vulnerability, adaptive capacity and underpinning values remains limited and poorly integrated with bio-physical studies. This limits the level of confidence in conclusions regarding future vulnerabilities and the feasibility and effectiveness of adaptation strategies. These multiple, persistent and structural uncertainties imply that, in most cases, adaptation requires an iterative risk management process. While decision-support frameworks are being developed, it remains unclear to what extent existing governance and institutional arrangements will be able to support more transformational responses, particularly where competing public and private interests and particularly vulnerable groups are involved. The enabling or constraining influences on adaptation from interactions among market forces, institutions, governance, policy and regulatory environments have only recently begun to attract research attention, mostly in Australia. Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climate change impacts, adaptation and mitigation responses in other world regions will affect Australasia, but our understanding of this remains very limited. Existing studies suggest that transboundary effects, mediated mostly via trade but potentially also migration, can be of similar if not larger scale than direct domestic impacts of climate change for economically important sectors such as agriculture and tourism. However, scenarios used in such studies tend to be highly simplified. Effective management of risks and opportunities in these sectors would benefit from better integration of relevant global scenarios of climatic and socio-economic changes into studies of local vulnerability and adaptation options. Frequently Asked Questions FAQ 25-1: How can we adapt to climate change if projected future changes remain uncertain? [to be inserted at end of Section 25.4.2] Many existing climate change impact assessments in Australia and New Zealand focus on the distant future (2050 to 2100). When contrasted with more near-term non-climate pressures, the inevitable uncertainty of distant climate impacts can impede effective adaptation. Emerging best practice in Australasia recognises this challenge and instead focuses on those decisions that can and will be made in the near future in any case, along with the lifetime of those decisions, and the risk from climate change during that lifetime. Thus, for example, the choice of next year s annual crop, even though it is greatly affected by climate, only matters for a year or two and can be adjusted relatively quickly. Even land-use change among cropping, grazing and forestry industries has demonstrated significant flexibility in Australasia over the space of a decade. When the adaptation challenge is reframed as implications for near-term decisions, uncertainty about the distant future becomes less problematic and adaptation responses can be better integrated into existing decision-making processes and early warning systems. Some decisions, such as those about long-lived infrastructure and spatial planning and of a public good nature, must take a long-term view and deal with significant uncertainties and trade-offs between short- and long-term goals and values. Even then, widely used techniques can help reduce challenges for decision-making including the precautionary principle , real options , adaptive management , no regrets strategies , or risk hedging . These can be matched to the type of uncertainty but depend on a regulatory framework and institutions that can support such approaches, including the capacity of practitioners to implement them robustly. Adaptation is not a one-off action but will take place along an evolving pathway, in which decisions will be revisited repeatedly as the future unfolds and more information comes to hand (see Figure 25-3). Although this creates learning opportunities, successive short-term decisions need to be monitored to avoid unwittingly creating an adaptation path that is not sustainable as climate change continues, or which would cope only with a limited sub-set of possible climate futures. This is sometimes referred to as maladaptation. Changing pathways for example, shifting from on-going coastal protection to gradual retreat from the most exposed areas can be challenging and may require new types of interactions among governments, industry and communities. [INSERT FIGURE 25-3 HERE Adaptation as an iterative risk management process. Individual adaptation decisions comprise well known aspects of risk assessment and management (top left panel). Each decision occurs within and exerts its own sphere of influence, determined by the lead- and consequence time of the decision, and the broader regulatory and societal influences on the decision (top right panel). A sequence of adaptation decisions creates an adaptation pathway (bottom panel). There is no single correct adaptation pathway, although some decisions, and sequences of decisions, are more likely to result in long-term maladaptive outcomes than others, but the judgment of outcomes depends strongly on societal values, expectations and goals.] FAQ 25-2: What are the key risks from climate change to Australia and New Zealand? [to be inserted at end of Section 25.10.3] Our assessment identifies eight key regional risks from climate change. Some impacts, especially on ecosystems, are by now difficult to avoid entirely. Coral reef systems have a limited ability to adapt naturally to further warming and an increasingly acidic ocean. Similarly, the habitat for some mountain or high elevation ecosystems and their associated species is shrinking inexorably with rising temperatures. This implies substantial impacts and some losses even under scenarios of limited warming. Other risks, however, can be reduced substantially by adaptation, combined with globally effective mitigation. These include potential flood damages from more extreme rainfall in Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 most parts of Australia and New Zealand; constraints on water resources from reducing rainfall in southern Australia; increased health risks and infrastructure damages from heat waves in Australia; and, increased economic losses, risks to human life and ecosystem damage from wildfires in southern Australia and many parts of New Zealand. A third set of risks is particularly challenging to manage robustly because the severity of potential impacts varies widely across the range of climate projections, even for a given temperature increase. These concern damages to coastal infrastructure and low-lying ecosystems from continuing sea level rise, where damages would be widespread if sea level turns out to be at the upper end of current scenarios; and, threats to agricultural production in both far south-eastern and far south-western Australia, which would affect ecosystems and rural communities severely at the dry end of projected rainfall changes. Even though some of these key risks are more likely to materialise than others, and they differ in the extent that they can be managed by adaptation and mitigation, they all warrant attention from a risk management perspective, given their potential major consequences for the region. References ABARES, 2011: Australian mineral statistics 2011, March Quarter 2011. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra, 42 pp. ABARES, 2012: Australia s Forests at a Glance 2012. Australian Bureau of Agricultural and Resource Economics and Sciences, Canberra, 83+vii pp. Abbs, D.J., A. Rafter, 2009: Impact of Climate Variability and Climate Change on Rainfall Extremes in Western Sydney and Surrounding Areas: Component 4 - Dynamical Downscaling. Report to the Sydney Metro Catchment Management Authority and Partners. CSIRO, Aspendale, 83 pp. Abbs, D.J., 2012: The impact of climate change on the climatology of tropical cyclones in the Australian region. Climate Adaptation Flagship Working Paper No 11. CSIRO, Aspendale, 24 pp. Abdo, D.A., D.A. Bellchambers, L.M. Evans, S.N. Evans, 2012: Turning up the heat: increasing temperature and coral bleaching at the high latitude coral reefs of the Houtman Abrolhos Islands. PLOS ONE, 7(8), e43878. Abel, N., R. Gorddard, B. Harman, A. Leitch, J. Langridge, A. Ryan, S. Heyenga, 2011: Sea level rise, coastal development and planned retreat: analytical framework, governance principles and an Australian case study. Environmental Science & Policy, 14(3), 279-288. ABS, 2008: Population Projections Australia, 2006 to 2101. Catalogue No. 3222. Australian Bureau of Statistics, Canberra, 100 pp. ABS, 2009: Experimental estimates and projections, Aboriginal and Torres Strait Islander Australians, Catalogue 3238.0. Australian Bureau of Statistics, Canberra, 76 pp. ABS, 2010a: Tourism satellite account 2009-10. Catalogue 5249.0. Australian Bureau of Statistics, Canberra, 50 pp. ABS, 2010b: Population characteristics, Aboriginal and Tores Trait Islander Australians. Catalogue No. 4713.0. Australian Bureau of Statistics, Canberra, 200 pp. ABS, 2010c: The Health and Welfare of Australia's Aboriginal and Torres Strait Islander peoples, 2010. Catalogue 4704.0. Australian Bureau of Statistics, Canberra, 107 pp. ABS, 2012a: Water Account Australia 2009-10. Catalogue 4610.0. Australian Bureau of Statistics, Canberra, 93 pp. ABS, 2012b: Gross Value of Irrigated Agricultural Production, 2010-11. Catalogue No. 4610.0.55.008 (online at www.abs.gov.au). Australian Bureau of Statistics, Canberra. ABS, 2012c: Yearbook Australia 2012. Catalogue No. 1301.0. Australian Bureau of Statistics, Canberra, 885 pp. ABS, 2012d: Australian National Accounts: National Income, Expenditure and Product, Jun 2012. Table 5206.0. Australian Bureau of Statistics, Canberra, 81 pp. ABS, 2013: Australian Demographic Statistics. Australian Bureau of Statistics, Canberra, 108 pp. Ackerley, D., R. Bell, B. Mullan, H. McMillan, 2013: Estimation of regional departures from global-average sea- level rise around New Zealand from AOGCM simulations. Weather and Climate, 33, 2-22. ACT Government, 2012: AP2 A New Climate Change Strategy and Action Plan for the Australian Capital Territory. Environment and Sustainable Development Directorate, ACT Government, Canberra, 94 pp. Adams, M., P. Attiwill, 2011: Burning Issues: Sustainability and Management of Australia's Southern Forests. CSIRO Publishing, Collingwood, Vic, 144 pp. Adams-Hosking, C., H.S. Grantham, J.R. Rhodes, C. McAlpine, P.T. Moss, 2011: Modelling climate-change- induced shifts in the distribution of the koala. Wildlife Research, 38(2), 122-130. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Agho, K., G. Stevens, M. Taylor, M. Barr, B. Raphael, 2010: Population risk perceptions of global warming in Australia. Environmental Research, 110(8), 756-763. AGO, 2006: Climate Change Impacts and Risk Management: A Guide for Business and Government. Australian Greenhouse Office, Canberra, 73 pp. Akompad, D.A., P. Bi, S. Williams, J. Grant, I.A. Walker, M. Augoustinous, 2013: Heat waves and climate change: applying the health belief model to identify predicotrs of risk preception and adapative behaviours in Adelaide, Australia. International Journal of Environmental Reserach and Public Health, 10, 2164-2184. Alexander, L.V., P. Hope, D. Collins, B. Trewin, A. Lynch, N. Nicholls, 2007: Trends in Australia's climate means and extremes: a global context. Australian Meteorological Magazine, 56(1), 1-18. Alexander, L.V., J.M. Arblaster, 2009: Assessing trends in observed and modelled climate extremes over Australia in relation to future projections. International Journal of Climatology, 29(3), 417-435. Alexander, L.V., X.L. Wang, H. Wan, B. Trewin, 2011: Significant decline in storminess over southeast Australia since the late 19th century. Australian Meteorological and Oceanographic Journal, 61, 23-30. Alexander, K.S., A. Ryan, T.G. Measham, 2012: Managed retreat of coastal communities: understanding responses to projected sea level rise. Journal of Environmental Planning and Management, 55(4), 409-433. Allen, C.D., A.K. Macalady, H. Chenchouni, D. Bachelet, N. McDowell, M. Vennetier, T. Kitzberger, A. Rigling, D.D. Breshears, E.H. Hogg, P. Gonzalez, R. Fensham, Z. Zhang, J. Castro, N. Demidova, J.H. Lim, G. Allard, S.W. Running, A. Semerci, N. Cobb, 2010: A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. Forest Ecology and Management, 259(4), 660-684. Allen, J.T., D.J. Karoly, 2013: A climatology of Australian severe thunderstorm environments 1979 2011: inter- annual variability and ENSO influence. International Journal of Climatology, advance online, DOI: 10.1002/joc.3667. Alston, M., 2007: Gender and climate change: variable adaptations of women and men. Just Policy, 46, 29-35. Alston, M., 2010: Gender and climate change in Australia. Journal of Sociology, 47(1), 53-70. Alston, M., 2012: Rural male suicide in Australia. Social Science & Medicine, 74(4), 512-522. Altman, J., G. Buchanan, L. Larsen, 2007: The environmental significance of the Indigenous estate: natural resource management as economic development in remote Australia. CAEPR Discussion Paper No. 286/2007. Centre for Aboriginal Economic Policy Research, ANU, Canberra, 65 pp. Anderson, I., S. Crengle, M. Leialoha Kamaka, T.-H. Chen, N. Palafox, L. Jackson-Pulver, 2006: Indigenous health in Australia, New Zealand, and the Pacific. The Lancet, 367(9524), 1775-1785. Anderson, B., W. Lawson, I. Owens, 2008: Response of Franz Josef Glacier Ka Roimata o Hine Hukatere to climate change. Global and Planetary Change, 63(1), 23-30. Anthony, K.R.N., D.I. Kline, G. Diaz-Pulido, S. Dove, O. Hoegh-Guldberg, 2008: Ocean acidification causes bleaching and productivity loss in coral reef builders. Proceedings of the National Academy of Sciences, 105(45), 17442-17446. ANU, 2009: Implications of climate change for Australia s World Heritage properties: A preliminary assessment. Report by the Fenner School of Environment and Society, Australian National University. Department of Climate Change, Department of the Environment, Water, Heritage and the Arts, Canberra, 207 pp. Arkema, K.K., G. Guannel, G. Verutes, S.A. Wood, A. Guerry, M. Ruckelshaus, P. Kareiva, M. Lacayo, J.M. Silver, 2013: Coastal habitats shield people and property from sea-level rise and storms. Nature Climate Change, advance online, DOI: 10.1038/nclimate1944. Ashworth, P., T. Jeanneret, J. Gardner, H. Shaw, 2011: Communication and climate change: What the Australian public thinks. Report EP112769. CSIRO, Canberra, 68 pp. Asseng, S., F. Ewert, C. Rosenzweig, J.W. Jones, J.L. Hatfield, A.C. Ruane, K.J. Boote, P.J. Thorburn, R.P. Rotter, D. Cammarano, N. Brisson, B. Basso, P. Martre, P.K. Aggarwal, C. Angulo, P. Bertuzzi, C. Biernath, A.J. Challinor, J. Doltra, S. Gayler, R. Goldberg, R. Grant, L. Heng, J. Hooker, L.A. Hunt, J. Ingwersen, R.C. Izaurralde, K.C. Kersebaum, C. Muller, S. Naresh Kumar, C. Nendel, G. O/'Leary, J.E. Olesen, T.M. Osborne, T. Palosuo, E. Priesack, D. Ripoche, M.A. Semenov, I. Shcherbak, P. Steduto, C. Stockle, P. Stratonovitch, T. Streck, I. Supit, F. Tao, M. Travasso, K. Waha, D. Wallach, J.W. White, J.R. Williams, J. Wolf, 2013: Uncertainty in simulating wheat yields under climate change. Nature Climate Change, 3(9), 827-832. Aström, C., J. Rocklöv, S. Hales, A. Béguin, V. Louis, R. Sauerborn, 2012: Potential Distribution of Dengue Fever Under Scenarios of Climate Change and Economic Development. EcoHealth, 9(4), 448-454. ATSE, 2008: Assessment of impacts of climate change on Australia's physical infrastructure. The Australian Academy of Technological Sciences and Engineering (ATSE), Parkville, Vic, 71 pp. Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 ATSISJC, 2009: Native Title Report 2008. Aboriginal and Torres Strait Islander Social Justice Commissioner, Australian Human Rights Commission, Sydney, 422 pp. Auckland Council, 2012: The Auckland Plan. Auckland Council, Auckland, 373 pp. August, S.M., B.J. Hicks, 2008: Water temperature and upstream migration of glass eels in New Zealand: implications of climate change. Environmental Biology of Fishes, 81(2), 195-205. Australian Government, 2012: State of Australian Cities 2012. Department of Infrastructure and Transport, Australian Government, Canberra, 365 pp. Australian Treasury, 2010: Australia to 2050: future challenges. The 2010 Intergenerational Report. Australian Treasury, Canberra, 164 pp. Bacon, W., 2011: A sceptical climate. Media coverage of climate change in Australia. Part 1 - Climate change policy. Australian Centre for Independent Journalism, University of Technology, Sydney, 70 pp. Baisden, W.T., E.D. Keller, L. Timar, D. Smeaton, A. Clark, A. Ausseil, W. Power, W. Zhang, 2010: New Zealand s Pasture Production in 2020 and 2050. Consultancy Report 2010/154 by GNS Sciences for the Ministry of Agriculture and Forestry. Ministry for Primary Industries, Wellington, New Zealand, 68 pp. Ballinger, J., B. Jackson, A. Reisinger, K. Stokes, 2011: The potential effects of climate change on flood frequency in the Hutt River. School of Geography, Environment and Earth Sciences, Victoria University of Wellington, Wellington, NZ, 40 pp. Balston, J.M., 2012: Guidelines for developing a climate change adaptation plan and undertaking an integrated climate change vulnerability assessment. Local Government Association of South Australia, Adelaide, 61 pp. Balston, J.M., J. Kellett, G. Wells, S. Li, A. Gray, I. Iankov, 2012: Quantifying the costs of climate change on local government assets. National Climate Change Adaptation Research Facility, Gold Coast, Qld, 219 pp. Bambrick, H.J., K. Dear, R. Woodruff, I. Hanigan, A.M. Michael, 2008: The impacts of climate change on three health outcomes. Paper prepared for Garnaut Climate Change Review. Garnaut Climate Change Review, Canberra, 47 pp. Banfai, D.S., D. Bowman, 2007: Drivers of rain-forest boundary dynamics in Kakadu National Park, northern Australia: a field assessment. Journal of Tropical Ecology, 23, 73-86. Banks, S.C., S.D. Ling, C.R. Johnson, M.P. Piggott, J.E. Williamson, L.B. Beheregaray, 2010: Genetic structure of a recent climate change-driven range extension. Molecular Ecology, 19(10), 2011-2024. Barnett, J., 2009: Climate change and human security in the Pacific Islands: the potential for and limits to adaptation. In: Climate change and security. Planning for the future [Boston, J., Nel, P., Righarts, M. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, NZ, pp. 59-70. Barnett, J., S. O'Neill, 2010: Maladaptation. Global Environmental Change, 20(2), 211-213. Barnett, J., E. Waters, S. Pendergast, A. Puleston, 2013: Barriers to adaptation to sea-level rise. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 85 pp. Barton, A.B., J.R. Argue, 2009: Integrated urban water management for residential areas: a reuse model. Water Science and Technology, 60(3), 813-823. Basher, L., S. Elliot, A. Hughes, A. Tait, M. Page, B. Rosser, I. McIvor, G. Douglas, H. Jones, 2012: Impacts of climate change on erosion and erosion control methods A critical review. MPI Technical Paper No: 2012/45. Ministry for Primary Industries, Wellington, New Zealand, 208 pp. Bates, B.C., P. Hope, B. Ryan, I. Smith, S. Charles, 2008: Key findings from the Indian Ocean Climate Initiative and their impact on policy development in Australia. Climatic Change, 89(3-4), 339-354. Bates, B.C., R.E. Chandler, S.P. Charles, E.P. Campbell, 2010: Assessment of apparent nonstationarity in time series of annual inflow, daily precipitation, and atmospheric circulation indices: A case study from southwest Western Australia. Water Resources Research, 46, W00H02. Bates, B.C., S. Bunn, P. Baker, M. Cox, A. Hopkins, B. Humphreys, S. Lakes, G. Willgoose, B. Young, 2011: National Climate Change Adaptation Research Plan for Freshwater Biodiversity. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 68 pp. Battaglene, S.C., C. Carter, A.J. Hobday, V. Lyne, B. Nowak, 2008: 'Scoping Study into Adaptation of the Tasmanian Salmonid Aquaculture Industry to Potential Impacts of Climate Change. National Agriculture & Climate Change Action Plan: Implementation Programme report. Tasmanian Aquaculture and Fisheries Institute, University of Tasmania, Hobart, 84 pp. Battaglia, M., J. Bruce, C. Brack, T. Baker, 2009: Climate Change and Australia's Plantation Estate: Analysis of Vulnerability and Preliminary Investigation of Adaptation Options. Forest & Wood Products Australia Limited, Melbourne, Vic, 130 pp. Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Baum, S., S. Horton, D.L. Choy, B. Gleeson, 2009: Climate change, health impacts and urban adaptability: case study of Gold Coast City. Research Monograph 11. Griffith University, Urban Research Program, Brisbane, Qld, 76 pp. Baynes, T., A. Herr, A. Langston, H. Schandl, 2012: Coastal Climate Risk Project Milestone 1 Final report. Department of Climate Change and Energy Efficiency, Canberra, 104 pp. Becken, S., 2011: A Critical Review of Tourism and Oil. Annals of Tourism Research, 38(2), 359-379. Becken, S., R. Clapcott, 2011: National tourism policy for climate change. Journal of Policy Research in Tourism, Leisure and Events, 3(1), 1-17. Bedford, R., C. Bedford, 2010: International Migration and Climate Change: A Post-Copenhagen Perspective on Options for Kiribati and Tuvalu. In: Climate change and Migration: South Pacific perspectives [Burson, B. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, pp. 89-134. Bedford, R., J. Campbell, 2013: Migration and climate change in Oceania. In: People on the move in a changing climate [Piguet, E., Laczko, F. (eds.)]. Springer, Dordrecht (ISBN 978-94-007-6984-7). Beebe, N.W., R.D. Cooper, P. Mottram, A.W. Sweenet, 2009: Australia's dengue risk driven by human adaptation to climate change. PLOS Neglected Tropical Diseases, 3(5), e429. Beentjes, M.P., J.A. Renwick, 2001: The relationship between red cod, Pseudophycis bachus, recruitment and environmental variables in New Zealand. Environmental Biology of Fishes, 61(3), 315-328. Beggs, P.J., C.M. Bennett, 2011: Climate change, aeroallergens, natural particulates and human health in Australia: state of science and policy. Asia Pacific Journal of Public Health, 23, 46S-53S. Béguin, A., S. Hales, J. Rocklöv, C. Aström, V.R. Louis, R. Sauerborn, 2011: The opposing effects of climate change and socio-economic development on the global distribution of malaria. Global Environmental Change, 21(4), 1209-1214. Bell, M., G. Blick, O. Parkyn, P. Rodway, P. Vowles, 2010: Challenges and Choices: Modelling New Zealand s Long-term Fiscal Position. Working Paper 10/01. New Zealand Treasury, Wellington, NZ, 85 pp. Bennett, C.M., K.B.G. Dear, A.J. McMichael, 2013: Shifts in the seasonal dsitribution of deaths in Australia, 1968- 2007. International Journal of Biometeorology, advance online, DOI: 10.1007/s00484-013-0663-x. Bergin, A., J. Townsend, 2007: A change in climate for the Australian defence force. Special Report 2007-7. Australian Strategic Policy Institute, Canberra, 11 pp. Berry, S., J. Vella, 2010: Planning Controls and Property Rights Striking the Balance. Roadshow 2010. Resource Management Law Association, Auckland, 66 pp. Berry, H.L., J.R.A. Butler, C.P. Burgess, U.G. King, K. Tsey, Y.L. Cadet-James, C.W. Rigby, B. Raphael, 2010: Mind, body, spirit: co-benefits for mental health from climate change adaptation and caring for country in remote Aboriginal Australian communities. NSW Public Health Bulletin, 21(5-6), 139-45. Betts, R.A., O. Boucher, M. Collins, P.M. Cox, P.D. Falloon, N. Gedney, D.L. Hemming, C. Huntingford, C.D. Jones, D.M.H. Sexton, M.J. Webb, 2007: Projected increase in continental runoff due to plant responses to increasing carbon dioxide. Nature, 448, 1037-1041. Bevin, S., 2007: Economic Impact of the 2007 East Coast Drought on the Sheep and Beef Sector. Report by Sean Bevin, Economic Solutions LTD, Napier. Ministry of Agriculture and Forestry, Wellington, 20 pp. Bi, P., K.A. Parton, 2008: Effect of climate change on Australian rural and remote regions: What do we know and what do we need to know? Australian Journal of Rural Health, 16(1), 2-4. Bicknell, S., P. McManus, 2006: The Canary in the Coalmine: Australian Ski Resorts and their Response to Climate Change. Geographical Research, 44(4), 386-400. Bird, D.W., R.B. Bird, C.H. Parker, 2005: Aboriginal burning regimes and hunting strategies in Australia's western desert. Human ecology, 33(4), 443-464. Blanchette, C.A., E.A. Wieters, B.R. Broitman, B.P. Kinlan, D.R. Schiel, 2009: Trophic structure and diversity in rocky intertidal upwelling ecosystems: A comparison of community patterns across California, Chile, South Africa and New Zealand. Progress in Oceanography, 83(1-4), 107-116. BMT WBM, 2011: Kakadu: Vulnerability to climate change impacts. Department of Climate Change and Energy Efficiency, Canberra, 229 pp. Bohensky, E., J.R.A. Butler, R. Costanza, I. Bohnet, A. Delisle, K. Fabricius, M. Gooch, I. Kubiszewski, G. Lukacs, P. Pert, E. Wolanski, 2011: Future makers or future takers? A scenario analysis of climate change and the Great Barrier Reef. Global Environmental Change, 21(3), 876-893. BoM, 2009: The exceptional January-February 2009 Heatwave in South-Eastern Australia. Special Climate Statement 17. National Climate Centre, Bureau of Meteorology, Canberra, 11 pp. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 BoM, 2013: Special Climate Statement 43 extreme heat in January 2013. Bureau of Meteorology, Canberra, 23 pp. Bond, N., P. Lake, A. Arthington, 2008: The impacts of drought on freshwater ecosystems: an Australian perspective. Hydrobiologia, 600(1), 3-16. Bond, N., J. Thomson, P. Reich, J. Stein, 2011: Using species distribution models to infer potential climate change- induced range shifts of freshwater fish in south-eastern Australia. Marine and Freshwater Research, 62(9), 1043-1061. Booth, T.H., M.U.F. Kirschbaum, M. Battaglia, 2010: Forestry. In: Adapting Agriculture to Climate Change. Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 137-152. Booth, K., S. Williams, 2012: Is insurance an under-utilised mechanism in climate change adaptation? The case of bushfire management in Tasmania. Australian Journal of Emergency Management, 27(4), 38-45. Botterill, L.C., S. Dovers, 2013: Drought and water in the Murray-Darling Basin: from disaster policy to adaptation. In: Natural Disasters and Adaptation to Climate Change [Boulter, S., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 100-114. Bourdôt, G.W., S.L. Lamoureaux, M.S. Watt, L.K. Manning, D.J. Kriticos, 2012: The potential global distribution of the invasive weed Nassella neesiana under current and future climates. Biological Invasions, 14(8), 1545- 1556. Bowman, D., B.P. Murphy, D.S. Banfai, 2010: Has global environmental change caused monsoon rainforests to expand in the Australian monsoon tropics? Landscape Ecology, 25(8), 1247-1260. Bradshaw, C.J.A., X. Giam, N.S. Sodhi, 2010: Evaluating the Relative Environmental Impact of Countries. Plos One, 5(5), e10440. Bradshaw, C.J.A., D.M.J.S. Bowman, N.R. Bond, B.P. Murphy, A.D. Moore, D.A. Fordham, R. Thackway, M.J. Lawes, H. McCallum, S.D. Gregory, R.C. Dalal, M.M. Boer, A.J.J. Lynch, R.A. Bradstock, B.W. Brook, B.K. Henry, L.P. Hunt, D.O. Fisher, D. Hunter, C.N. Johnson, D.A. Keith, E.C. Lefroy, T.D. Penman, W. Meyer, J.R. Thomson, C.M. Thornton, J. Van Der Wal, D. Williams, L. Keniger, A. Specht, 2013: Brave new green world consequences of a carbon economy for the conservation of Australian biodiversity. Biological Conservation, 161, 71-90. Bradstock, R.A., 2010: A biogeographic model of fire regimes in Australia: current and future implications. Global Ecology and Biogeography, 19(2), 145-158. BRANZ, 2007: An Assessment of the Need to Adapt Buildings for the Unavoidable Consequences of Climate Change. Report by the Building Research Association of New Zealand. Australian Greenhouse Office and Department of the Environment and Water Resources, Canberra, 81 pp. Bright, J., H. Rutter, J. Dommisse, R. Woods, A. Tait, B. Mullan, J. Hendrix, J. Diettrich, 2008: Projected Effects of Climate Change on Water Supply Reliability in mid-Canterbury. Report C08120/1 by Aqualine Research Ltd. Ministry of Agriculture and Forestry, Wellington, 45 pp. Britton, R., 2010: Coastal adaptation to climate change. Report on local government planning practice and limitations to adaptation. National Institute of Water and Atmospheric Research (NIWA), Wellington, 51 pp. Britton, E., S. Hales, K. Venugopal, M.G. Baker, 2010a: Positive association between ambient temperature and salmonellosis notifications in New Zealand, 1965-2006. Australian and New Zealand Journal of Public Health, 34(2), 126-129. Britton, E., S. Hales, K. Venugopal, M.G. Baker, 2010b: The impact of climate variability and change on cryptosporidiosis and giardiasis rates in New Zealand. Journal of Water and Health, 8(3), 561-571. Britton, R., J. Dahm, H. Rouse, T. Hume, R. Bell, P. Blackett, 2011: Coastal Adaptation to Climate Change: Pathways to Change. NIWA, Auckland, 106 pp. Brockerhoff, E.G., H. Jactel, J.A. Parrotta, C.P. Quine, J. Sayer, 2008: Plantation Forests and Biodiversity: Oxymoron or Opportunity? Biodiversity Conservation, 17, 925 951. Brown, A.E., L. Zhang, T.A. McMahon, A.W. Western, R.A. Vertessy, 2005: A review of paired catchment studies for determining changes in water yield resulting from alterations in vegetation. Journal of Hydrology, 310, 28- 61. Brown, R.R., M. Farrelly, 2009: Challenges ahead: social and institutional factors influencing sustainable urban stormwater management in Australia. Water Science and Technology, 59(4), 653-660. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Brown, R.R., M. Farrelly, N. Keath, 2009: Practitioner Perceptions of Social and Institutional Barriers to Advancing a Diverse Water Source Approach in Australia. International Journal of Water Resources Development, 25(1), 15-28. Brunckhorst, D., I. Reeve, P. Morley, M. Coleman, E. Barclay, J. McNeill, R. Stayner, R. Glencross-Grant, J. Thompson, L. Thompson, 2011: Hunter & Central Coasts New South Wales Vulnerability to climate change impacts. Department of Climate Change and Energy Efficiency, Canberra, 149 pp. Bruno, J.F., E.R. Selig, K.S. Casey, C.A. Page, B.L. Willis, C.D. Harvell, H. Sweatman, A.M. Melendy, 2007: Thermal stress and coral cover as drivers of coral disease outbreaks. Plos Biology, 5(6), 1220-1227. Bryan, B.A., D. King, E.L. Wang, 2010: Biofuels agriculture: landscape-scale trade-offs between fuel, economics, carbon, energy, food, and fiber. Global Change Biology Bioenergy, 2(6), 330-345. BSC, 2010: Draft Coastal Zone Management Plan for Byron Shire Coastline. Part A: The Plan. Byron Shire Council, Mullumbimby, NSW, 413 pp. Buckley, R., 2008: Misperceptions of Climate Change Damage Coastal Tourism: Case Study of Byron Bay, Australia. Tourism Review International, 12, 71-88. Burbidge, A.A., M. Byrne, D. Coates, S.T. Garnett, S. Harris, M.W. Hayward, T.G. martin, E. McDonald-Madden, N.J. Mitchell, S. Nally, S.A. Setterfield, 2011: Is Australia ready for assisted colonization? Policy changes required to facilitate translocations under climate change. Pacific Conservation bBiology, 17(3), 259-269. Burgess, C.P., F.H. Johnston, H. Berry, J. McDonnell, D. Yibarbuk, C. Gunabarra, A. Mileran, R. Bailie, 2009: Healthy country healthy people: Superior Indigenous health outcomes are associated with 'caring for country'. Medical Journal of Australia, 190(10), 567-572. Burgette, R.J.W., C.S, J.A. Church, N.J. White, P. Tregoning, R. Coleman, 2013: Characterizing and minimizing the effects of noise in tide gauge time series: relative and geocentric sea level rise around Australia. Geophysical Journal International, advance online, DOI: 10.1093/gji/ggt131. Burroughs, 2010: Sustainable housing for Indigenous populations in remote regions: A case study from Nguiu, Northern Australia. In: Keynote Lectures: CESB 10: Central Europe towards Sustainable Building. From Theory to Practice, Czech Technical University, Prague, Czech Republic. pp 8. Burton, P., J. Mustelin, 2013: Planning for climate change: is greater public participation the key to success? Urban Policy and Research, advance online, DOI: 1080/08111146.2013.778196. Butcher, 2009: Regional and national impacts of the 2007 2009 drought. Report by Butcher Partners Ltd. Ministry of Agriculture and Forestry, Wellington, 33 pp. Buultjens, J., D. Gale, N.E. White, 2010: Synergies between Australian Indigenous tourism and ecotourism: possibilities and problems for future development. Journal of Sustainable Tourism, 18(4), 497-513. Byron, N., 2011: What Can the Murray-Darling Basin Plan Achieve? Will it be enough? In: Basin futures: water reform in the Murray-Darling Basin [Grafton, Q., Connell, D. (eds.)]. ANU E-Press, Canberra, pp. 385-398. Cai, W., T. Cowan, 2006: SAM and regional rainfall in IPCC AR4 models: Can anthropogenic forcing account for southwest Western Australian winter rainfall reduction? Geophysical Research Letters, 33(24), L24708. Cai, W., T. Cowan, 2008: Evidence of impacts from rising temperature on inflows to the Murray-Darling Basin. Geophysical Research Letters, 35, L07701. Cai, W., T. Cowan, M. Raupach, 2009a: Positive Indian Ocean Dipole events precondition southeast Australia bushfires. Geophysical Research Letters, 36(19), L19710. Cai, W., A. Sullivan, T. Cowan, 2009b: Climate change contributes to more frequent consecutive positive Indian Ocean Dipole events. Geophysical Research Letters, 36(23), L23704. Cai, W., P. van Rensch, T. Cowan, 2011: Influence of Global-Scale Variability on the Subtropical Ridge over Southeast Australia. Journal of Climate, 24(23), 6035-6053. Cai, W., P. van Rensch, T. Cowan, 2012: The 2011 southeast Queensland floods: a confirmation of a negative Pacific Decadal Oscillation phase? Geophysical Research Letters, 39(8), L08702. Calder, J.A., J.B. Kirkpatrick, 2008: Climate change and other factors influencing the decline of the Tasmanian cider gum (Eucalyptus gunnii). Australian Journal of Botany, 56(8), 684-692. Callaghan, J., S.B. Power, 2011: Variability and decline in the number of severe tropical cyclones making landfall over eastern Australia since the late nineteenth century. Climate Dynamics, 37(3-4), 647-662. Cameron, P.A., B. Mitra, M. Fitzgerald, C.D. Scheinkestel, A. Stripp, C. Batey, L. Niggemeyer, M. Truesdale, P. Holman, R. Mehra, J. Wasiak, H. Cleland, 2009: Black Saturday: the immediate impact of the February 2009 bushfires in Victoria, Australia. The Medical Journal of Australia, 191, 11-16. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Campbell, A., 2008: Managing Australian Landscapes in a Changing Climate: A climate change primer for regional Natural Resource Management bodies. Department of Climate Change, Canberra, 47 pp. Campbell, D., M. Stafford Smith, J. Davies, P. Kuipers, J. Wakerman, M.J. McGregor, 2008: Responding to health impacts of climate change in the Australian desert. Rural and Remote Health, 8(3), 1008. Carey-Smith, T., S. Dean, J. Vial, C. Thompson, 2010: Changes in precipitation extremes for New Zealand: climate model predictions. Weather and Climate, 30, 23-48. Carr, M.K.V., J.W. Knox, 2011: The water relations and irrigation requirements of sugarcane (Saccharum officinarum): A review. Experimental Agriculture, 47(1), 1-25. Carswell, F., G. Harmsworth, R. Kirikiri, I. Turney, S. Kerr, 2002: A Framework For Engagement Of M ori Landowners In Carbon Farming Using Indigenous Forest Regeneration. Landcare Research Contract Report: LC0102/116. Landcare Research, Lincoln, 24 pp. Carter, T.R., 2013: Agricultural impacts: Multi-model yield projections. Nature Climate Change, 3(9), 784-786. Cary, G., D. Lindenmayer, S. Dovers (eds.), 2003: Australia Burning - Fire Ecology, Policy and Management Issues. CSIRO Publishing, Collingwood, Vic, 280 pp. CCC, 2010: Christchurch City Council climate smart strategy 2010-2025. Christchurch City Council, Christchurch, 26 pp. Chakraborty, S., J. Luck, G. Hollaway, G. Fitzgerald, N. White, 2011: Rust-proofing wheat for a changing climate. Euphytica, 179, 19-32. Chambers, L.E., G.M. Griffiths, 2008: The changing nature of temperature extremes in Australia and New Zealand. Australian Meteorological Magazine, 57, 13-55 Chambers, L.E., C.A. Devney, B.C. Congdon, N. Dunlop, E.J. Woehler, P. Dann, 2011: Observed and predicted effects of climate on Australian seabirds. Emu, 111(3), 235-251. Chambers, L.E., R. Altwegg, C. Barbraud, P. Barnard, L.J. Beaumont, R.J.M. Crawford, J.M. Durant, L. Hughes, M.R. Keatley, M. Low, L.P.C. Morellato, E.S. Poloczanska, V. Ruoppolo, R.E.T. Vanstreels, E.J. Woehler, A.C. Wolfaardt, 2013a: Phenological changes in the Southern Hemisphere. Plos One, advance online, DOI: 10.1371/journal.pone.0075514. Chambers, L.E., L. Beaumont, I. Hudson, 2013b: Continental scale analysis of bird migration timing: influences of climate and life history traits; a generalised mixture model clustering approach. Int. J. Biometeorology, advance online, DOI: 10.1007/s00484-013-0707-2. Chessman, B.C., 2009: Climatic changes and 13-year trends in stream macroinvertebrate assemblages in New South Wales, Australia. Global Change Biology, 15(11), 2791-2802. Cheung, W.W.L., J.J. Meeuwig, M. Feng, E. Harvey, V.W.Y. Lam, T. Langlois, D. Slawinski, C.J. Sun, D. Pauly, 2012: Climate-change induced tropicalisation of marine communities in Western Australia. Marine and Freshwater Research, 63(5), 415-427. Chiew, F., T.A. McMahon, 2002: Global ENSO-streamflow teleconnection, streamflow forecasting and interannual variability. Hydrological Sciences Journal-Journal Des Sciences Hydrologiques, 47(3), 505-522. Chiew, F., 2006: Estimation of rainfall elasticity of streamflow in Australia. Hydrological Sciences Journal-Journal Des Sciences Hydrologiques, 51(4), 613-625. Chiew, F., J. Teng, J. Vaze, D.A. Post, J.M. Perraud, D.G.C. Kirono, N.R. Viney, 2009: Estimating climate change impact on runoff across southeast Australia: Method, results, and implications of the modeling method. Water Resources Research, 45(10), W10414. Chiew, F., I.P. Prosser, 2011: Water and Climate. In: Water: Science and Solutions for Australia [Prosser, I. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 29-46. Chiew, F., W.J. Young, W. Cai, J. Teng, 2011: Current drought and future hydroclimate projections in southeast Australia and implications for water resources management. Stochastic Environmental Research and Risk Assessment, 25(4), 601-612. Chiew, F., N.J. Potter, J. Vaze, C. Petheram, L. Zhang, J. Teng, D.A. David, 2013: Observed hydrologic non- stationarity in far south-eastern Australia: implications and future modelling predictions. Stochastic Environmental Research and Risk Assessment, advance online, DOI: 10.1007/s00477-013-0755-5. Chinn, T.J., 2001: Distribution of the glacial water resources of New Zealand. Journal of Hydrology (NZ), 40(2), 139-187. Chinn, T.J., S. Winkler, M.J. Salinger, N. Haakensen, 2005: Recent glacier advances in Norway and New Zealand: a comparison of their glaciological and meteorological causes. Geografiska Annaler: Series A, Physical Geography, 87(1), 141-157. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chinn, T.J., B.B. Fitzharris, A. Willsman, M.J. Salinger, 2012: Annual ice volume changes 1976 2008 for the New Zealand Southern Alps. Global and Planetary Change, 92 93, 105-118. Chiswell, S.M., J.D. Booth, 2005: Distribution of mid- and late-stage Jasus edwardsii phyllosomas: implications for larval recruitment processes. New Zealand Journal of Marine and Freshwater Research, 39(5), 1157-1170. City of Adelaide, 2012: The City of Adelaide Strategic Plan 2012-2016. Adelaide City Council, Adelaide, 30 pp. City of Melbourne, 2009: City of Melbourne Climate Change Adaptation Strategy. Prepared by Maunsell Australia. City of Melbourne, Melbourne, 128 pp. City of Port Phillip, 2010: Greening Port Phillip, an urban forest approach. Port Phillip City Council, St Kilda, Vic, 54 pp. Clark, A., A. Tait, 2008: Drought, Agricultural Production & Climate Change A Way Forward to a Better Understanding. NIWA Client Report WLG2008-33. National Institute of Water & Atmospheric Research Ltd, Wellington, 71 pp. Clark, A., B. Mullan, A. Porteous, 2011: Scenarios of regional drought under climate change. NIWA Client Report WLG2012-32. National Institute of Water and Atmospheric Research (NIWA), Wellington, 135 pp. Clark, A., R. Nottage, D. Hansford (eds.), 2012: Impacts of Climate Change on Land-based Sectors and Adaptation Options. Stakeholder Report. Ministry of Primary Industries, Wellington, 74 pp. Clarke, H.G., P.L. Smith, A.J. Pitman, 2011: Regional signatures of future fire weather over eastern Australia from global climate models. International Journal of Wildland Fire, 20(4), 550-562. Clarke, H.G., C. Lucas, P.L. Smith, 2012: Changes in Australian fire weather between 1973 and 2010. International Journal of Climatology, 33(4), 931 944. Clayton, H., S. Dovers, P. Harris, 2011: Natural resource management policy and planning in Australia. Synthesis, Literature Review and Workshop Report. HC Coombs Policy Forum-Fenner School of Environment and Society NRM initiative, The Australian National University, Canberra, 37 pp. Clothier, B., A. Hall, S. Green, 2012: Horticulture: adapting the horticultural and vegetable industries to climate change. In: Enhanced climate change impact and adaptation evaluation: A comprehensive analysis of New Zealand's land-based primary sectors [Clark, A., Nottage, R. (eds.)]. Ministry for Primary Industries, Wellington, pp. 237-292. Clucas, R., 2011: Long-term population trends of Sooty Shearwater (Puffinus griseus) revealed by hunt success. Ecological Applications, 21(4), 1308-1326. COAG, 2007: National climate change adaptation framework. Council of Australian Governments, Department of the Prime Minister and Cabinet, Canberra, 27 pp. Cobon, D.H., G.S. Stone, J.O. Carter, J.C. Scanlan, N.R. Toombs, X. Zhang, J. Willcocks, G.M. McKeon, 2009: The climate change risk management matrix for the grazing industry of northern Australia. The Rangeland Journal, 31(1), 31-49. Cock, M.J.W., J.C. van Lenteren, J. Brodeur, B.I.P. Barratt, F. Bigler, K. Bolckmans, F.L. Cônzoli, F. Haas, P.G. Mason, J.G.P. Parra, 2010: Do new access and benefit sharing procedures under the convention on biological diversity threaten the future of biological control? BioControl, 55, 199-218. Cocklin, C., J. Dibden, 2009: Systems in peril: Climate change, agriculture and biodiversity in Australia. IOP Conference Series: Earth and Environmental Science, 8(1), 012013. Collins, D., R. Kearns, 2010: "It's a gestalt experience": Landscape values and development pressure in Hawke's Bay, New Zealand. Geoforum, 41(3), 435-446. Connell, D., 2007: Water politics in the Murray-Darling Basin. The Federation Press, Sydney. Connell, D., Q. Grafton (eds.), 2011: Basin Futures: water reform in the Murray-Darling Basin. ANU E-Press, Canberra, 500 pp. Connor, J., J. Schwabe, D. King, K. D., M. Kirby, 2009: Impacts of climate change on lower Murray irrigation. Australian Journal of Agricultural and Resources Economics, 53, 437-456. Cooper, T.F., G. De 'ath, K.E. Fabricius, J.M. Lough, 2008: Declining coral calcification in massive Porites in two nearshore regions of the northern Great Barrier Reef. Global Change Biology, 14(3), 529-538. Cooper, T.F., R.A. O'Leary, J.M. Lough, 2012: Growth of Western Australian Corals in the Anthropocene. Science, 335(6068), 593-596. Corkhill, J.R., 2013: Claimed property right does not hold water. Australian Law Journal, 87, 49-57. Cornish, P.M., R.A. Vertessy, 2001: Forest age-induced changes in evapotranspiration and water yield in a eucalypt forest. Journal of Hydrology, 242(1-2), 43-63. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cottrell, B., C. Insley, R. Meade, J. West, 2004: Report of the Climate Change M ori Issues Group. New Zealand Climate Change Office, Wellington, New Zealand, 27 pp. Coutts, A., J. Beringer, N. Tapper, 2010: Changing Urban Climate and CO2 Emissions: Implications for the Development of Policies for Sustainable Cities. Urban Policy and Research, 28(1), 27-47. Crase, L., 2011: The Fallout to the Guide to the Proposed Basin Plan. Australian Journal of Public Administration, 70(1), 84-93. Crompton, R.P., K.J. McAneney, 2008: Normalised Australian insured losses from meteorological hazards: 1967 2006. Environmental Science & Policy, 11(5), 371-378. Crompton, R.P., K.J. McAneney, K. Chen, R.A. Pielke, K. Haynes, 2010: Influence of Location, Population, and Climate on Building Damage and Fatalities due to Australian Bushfire: 1925 2009. Weather, Climate, and Society, 2(4), 300-310. Crook, J.A., L.A. Jones, P.M. Forster, R. Crook, 2011: Climate change impacts on future photovoltaic and concentrated solar power energy output. Energy & Environmental Science, 4(9), 3101-3109. Crosbie, R.S., J.L. McCallum, G.R. Walker, F. Chiew, 2010: Modelling climate-change impacts on groundwater recharge in the Murray-Darling Basin, Australia. Hydrogeology Journal, 18(7), 1639-1656. Crosbie, R.S., T. Pickett, F.S. Mpelasoka, G. Hodfgson, S.P. Charles, O. Barron, 2012: An assessment of the climate change impacts on groundwater recharge at a continental scale using a probabilistic approach with an ensemble of GCMs. Climatic Change, 117(1-2), 41-53. Crossman, N.D., B.A. Bryan, D.M. Summers, 2012: Identifying priority areas for reducing species vulnerability to climate change. Diversity and Distributions, 18(1), 60-72. Cruz, F., A. Pitman, Y. Wang, 2010: Can the stomatal response to higher atmospheric carbon dioxide explain the unusual temperatures during the 2002 Murray-Darling Basin drought? Journal of Geophysical Research- Atmospheres, 115(D2), D02101. CSIRO, BoM, 2007: Climate change in Australia. CSIRO and Bureau of Meteorology, Melbourne, 140 pp. CSIRO, Maunsell Australia Pty Ltd, Phillips Fox, 2007: Infrastructure and climate change risk assessment for Victoria. Report prepared for the Victorian Government. CSIRO, Aspendale, 76 pp. CSIRO, 2008: Water availability in the Murray-Darling Basin. A report from CSIRO to the Australian Government. CSIRO, Collingwood, 68 pp. CSIRO, 2009: Surface Water Yields in South-West Western Australia. A report to the Australian government from the CSIRO South-West Western Australia Sustainable Yields Project. CSIRO, Black Mountain, ACT, 195 pp. CSIRO, 2010: Climate Variability and Change in South-Eastern Australia: A Synthesis of Findings from Phase 1 of the South Eastern Australian Climate Initiative (SEACI). CSIRO, Collingwood, 31 pp. CSIRO, 2011: CSIRO, submission No. 40 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. CSIRO, Canberra (online at www.pc.gov.au/projects/inquiry/climate-change- adaptation/submissions), 31 pp. CSIRO, 2012: Climate and water availability in south-eastern Australia: A synthesis of findings from Phase 2 of the South Eastern Australian Climate Initiative (SEACI). CSIRO, Melbourne, 41 pp. CSIRO, BoM, 2012: State of the Climate 2012. CSIRO, Canberra, 12 pp. Cullen, J.M., L.E. Chambers, P.C. Coutin, P. Dann, 2009: Predicting onset and success of breeding in little penguins Eudyptula minor from ocean temperatures. Marine Ecology-Progress Series, 378, 269-278. Cullen, B.R., R.J. Eckard, 2011: Impacts of future climate scenarios on the balance between productivity and total greenhouse gas emissions from pasture based dairy systems in south-eastern Australia. Animal Feed Science and Technology, 166 167(0), 721-735. DAFF, 2012: Australia's agriculture, fisheries and forestry at a glance 2012. Department of Agriculture, Fisheries and Forestry, Canberra, 140 pp. Dalton, S.J., S. Godwin, S.D.A. Smith, L. Pereg, 2010: Australian subtropical white syndrome: a transmissible, temperature-dependent coral disease. Marine and Freshwater Research, 61(3), 342-350. Darbyshire, R., L. Webb, I. Goodwin, E.W.R. Barlow, 2013: Evaluation of recent trends in Australian pome fruit spring phenology. International Journal of Biometeorology, 57(3), 409-421. Davies, P.M., 2010: Climate Change Implications for River Restoration in Global Biodiversity Hotspots. Restoration Ecology, 18(3), 261-268. Davies, J., D. Campbell, M. Campbell, J. Douglas, H. Hueneke, M. LaFlamme, D. Pearson, K. Preuss, J. Walker, F.J. Walsh, 2011: Attention to four key principles can promote health outcomes from desert Aboriginal land management. The Rangeland Journal, 33(4), 417-431. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Davis, J., A. Pavlova, R. Thompson, P. Sunnucks, 2013: Evolutionary refugia and ecological refuges: key concepts for conserving Austrlaian arid zone reshwater biodiversity under climate change. Global Change Biology, 19, 1970-1984. DCC, 2009: Climate change risks to Australia s coast: A first-pass national assessment. Department of Climate Change, Government of Australia, Canberra, Australia, 172 pp. DCCEE, 2010: Developing a National Coastal Adaptation Agenda: A report on the national climate change forum, Adelaide, 18-19 February 2010. Department of Climate Change and Energy Efficiency, Canberra, 48 pp. DCCEE, 2011: Climate Change Risks to Coastal Buildings and Infrastructure; a supplement to the first pass national assessment. Department of Climate Change and Energy Efficiency, Canberra, 16 pp. De'ath, G., J.M. Lough, K.E. Fabricius, 2009: Declining Coral Calcification on the Great Barrier Reef. Science, 323(5910), 116-119. De'ath, G., K.E. Fabricius, H. Sweatman, M. Puotinen, 2012: The 27-year decline of coral cover on the Great Barrier Reef and its causes. Proceedings of the National Academy of Sciences of the United States of America, 109(44), 17995-17999. Dean, S., P. Stott, 2009: The Effect of Local Circulation Variability on the Detection and Attribution of New Zealand Temperature Trends. Journal of Climate, 22(23), 6217-6229. Deo, R., J. Syktus, C. McAlpine, P. Lawrence, H. McGowan, S. Phinn, 2009: Impact of historical land cover change on daily indices of climate extremes including droughts in eastern Australia. Geophysical Research Letters, 36(8), L08705. Depczynski, M.R., J.P. Gilmour, T. Ridgway, H. Barnes, A.J. Heyward, T.H. Holmes, J.A.Y. Moore, B.T. Radford, D. Thomson, P. Tinkler, S.K. Wilson, 2013: Bleaching, coral mortality adn subsequent survivorship on a West Australian fringing reef. Coral Reefs, 32, 233-238. DERM, DIP, LGAQ, 2010: Increasing Queensland s resilience to inland ooding in a changing climate: Final report on the Inland Flooding Study. Joint report by Department of Environment and Resource Management, Department of Infradtructure and Planning, and Local Government Association of Queensland. State of Queensland, Brisbane, Qld, 16 pp. Derraik, J.G.B., D. Slaney, 2007: Anthropogenic Environmental Change, Mosquito-borne Diseases and Human Health in New Zealand. EcoHealth, 4(1), 72-81. Derraik, J.G.B., D. Slaney, E.R. Nye, P. Weinstein, 2010: Chikungunya Virus: A Novel and Potentially Serious Threat to New Zealand and the South Pacific Islands. American Journal of Tropical Medicine and Hygiene, 83(4), 755-759. Diamond, H.J., A.M. Lorrey, J.A. Renwick, 2012: A Southwest Pacific Tropical Cyclone Climatology and Linkages to the El Nino Southern Oscillation. Journal of Climate, 26(1), 3-25. Diaz-Pulido, G., L.J. McCook, S. Dove, R. Berkelmans, G. Roff, D.I. Kline, S. Weeks, R.D. Evans, D.H. Williamson, O. Hoegh-Guldberg, 2009: Doom and Boom on a Resilient Reef: Climate Change, Algal Overgrowth and Coral Recovery. Plos One, 4(4), e5239. DIICCSRT, 2013: Climate Adaptation Outlook. A Proposed National Adaptation Assessment Framework. Department of Industry, Innovation, Climate Change, Science, Research and Tertiary Education, Canberra, 121 pp. DNP, 2010: Kakadu National Park Climate Change Strategy 2010-2015. Director of National Parks, Department of the Environment, Water, Heritage and the Arts, Canberra, 15 pp. Dobes, L., 2010: Notes on applying real options to climate change adaptation measures, with examples from Vietnam. Crawford School, Australian National University, Canberra, 25 pp. Doherty, T.J., S. Clayton, 2011: The Psychological Impacts of Global Climate Change. American Psychologist, 66(4), 265 276. Donohue, R.J., T.R. McVicar, M.L. Roderick, 2009: Climate-related trends in Australian vegetation cover as inferred from satellite observations, 1981-2006. Global Change Biology, 15(4), 1025-1039. Dovers, S., 2009: Normalizing adaptation. Global Environmental Change, 19(1), 4-6. Dovers, S., 2013: How much adaptation: are existing policy and institutions enough? In: Climate Adaptation Futures [Palutikof, J., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 97-102. Driscoll, D.A., D.B. Lindenmayer, A.F. Bennett, M. Bode, R.A. Bradstock, G.J. Cary, M.F. Clarke, N. Dexter, R. Fensham, G. Friend, M. Gill, S. James, G. Kay, D.A. Keith, C. MacGregor, J. Russell-Smith, D. Salt, J.E.M. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Watson, R.J. Williams, A. York, 2010: Fire management for biodiversity conservation: Key research questions and our capacity to answer them. Biological Conservation, 143(9), 1928-1939. Drucker, A.G., G.P. Edwards, W.K. Saalfeld, 2010: Economics of camel control in central Australia. The Rangeland Journal, 32(1), 117-127. DSE, 2007: Our Water Our Future: The Next Stage of the Government's Water Plan. Department of Sustainability and Environment Victoria, Melbourne, 28 pp. DSE, 2011: Guidelines for the development of a water supply demand strategy. Department of Sustainability and Environment Victoria, Melbourne, 100 pp. DSE, 2013: Victorian Climate Change Adaptation Plan. Department of Sustainability and Environment, State Government of Victoria, Melbourne, 96 pp. DSEWPC, 2011: Regional Natural Resources Management Planning for Climate Change Fund (online at www.environment.gov.au/cleanenergyfuture/regional-fund/). Department of the Environment (formerly Department of Sustainability, Environment, Water, Population and Communities), Canberra. Duckett, P.E., P.D. Wilson, A.J. Stow, 2013: Keeping up with teh neighbours: using a genetic measurement of dispersal and species dsitribution modelling to assess impact of climate change on an Australian aid zone gecko (Gehyra variegata). Diversity and Distributions, 19, 964-976. Duncan, M., G. Smart, 2011: Tool 2.1.4: Inundation modelling of present day and future floods, in: Impacts of Climate Change on Urban Infrastructure & the Build Environment - A Toolbox. NIWA, Wellington, 13 pp. Dunlop, M., P.R. Brown, 2008: Implications of climate change for Australia s National Reserve System: a preliminary assessment. Department of Climate Change, Canberra, 188 pp. Dunlop, M., D.W. Hilbert, S. Ferrier, A. House, A. Liedloff, S.M. Prober, A. Smyth, T.G. Martin, T. Harwood, K.J. Williams, C. Fletcher, H. Murphy, 2012: The Implications of Climate Change for Biodiversity Conservation and the National Reserve System: Final Synthesis. A report prepared for the Department of Sustainability, Environment, Water, Population and Communities, and the Department of Climate Change and Energy Efficiency. CSIRO Climate Adaptation Flagship, Canberra, 80 pp. Dunn, M.R., R.J. Hurst, J.A. Renwick, R.I.C.C. Francis, J. Devine, A. McKenzie, 2009: Fish abundance and climate trends in New Zealand. New Zealand Aquatic Environment and Biodiversity Report No. 31. Ministry of Fisheries, Wellington, 75 pp. Dupont, A., G. Pearman, 2006: Heating up the planet: climate change and security. Lowy Institute for International Policy, Sydney, 143 pp. Dupont, A., 2008: The Strategic Implications of Climate Change. Survival: Global Politics and Strategy, 50(3), 29- 54. Dwyer, A., C. Zoppou, O. Nielsen, S. Day, S. Roberts, 2004: Quantifying social vulnerability: a methodology for identifying those at risk to natural hazards. Geoscience Australia, Canberra, 92 pp. ECAN, 2005: Regional Coastal Environment Plan for the Canterbury Region. Environment Canterbury, Christchurch, 279 pp. Eckard, R.J., B.R. Cullen, 2011: Impacts of future climate scenarios on nitrous oxide emissions from pasture based dairy systems in south eastern Australia. Animal Feed Science and Technology, 166 167(0), 736-748. Edge Environment, 2011: Edge Environment, submission No. 54 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. Productivity Commission, Canberra (online at www.pc.gov.au/projects/inquiry/climate-change-adaptation/submissions), 7 pp. Edwards, B., M. Gray, 2009: A Sunburnt Country: The Economic and Financial Impact of Drought on Rural and Regional Families in Australia in an Era of Climate Change. Journal of Labour Economics, 12(1), 109-131. Edwards, F., J. Dixon, S. Friel, G. Hall, K. Larsen, S. Lockie, B. Wood, M. Lawrence, I. Hanigan, A. Hogan, L. Hattersley, 2011: Climate Change Adaptation at the Intersection of Food and Health. Asia-Pacific Journal of Public Health, 23(2, Supplement), 91-104. Elith, J., M. Kearney, S. Phillips, 2010: The art of modelling range-shifting species. Methods in Ecology and Evolution, 1(4), 330-342. Ellemor, H., 2005: Reconsidering emergency management and Indigenous communities in Australia. Global Environmental Change. Part B: Environmental Hazards, 6(1), 1-7. Energy Futures Forum, 2006: The heat is on. The future of energy in Australia. CSIRO, Canberra, 120 pp. Evans, L.S., P. Fidelman, C. Hicks, C. Morgan, A.L. Perry, R. Tobin, 2011: Limits to climate change adaptation in the Great Barrier Reef: scoping ecological and social limits. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 72 pp. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Evans, J.P., I. Boyer-Souchet, 2012: Local sea surface temperatures add to extreme precipitation in northeast Australia during La Nina. Geophys. Res. Lett., 39(10), L10803. Fabricius, K.E., C. Langdon, S. Uthicke, C. Humphrey, S. Noonan, G. De'ath, R. Okazaki, N. Muehllehner, M.S. Glas, J.M. Lough, 2011: Losers and winners in coral reefs acclimatized to elevated carbon dioxide concentrations. Nature Climate Change, 1(3), 165-169. Farbotko, C., H. Lazrus, 2012: The first climate refugees? Contesting global narratives of climate change in Tuvalu. Global Environmental Change, 22(2), 382-390. Fawcett, R.J.B., B.C. Trewin, K. Braganza, R.J. Smalley, B. Jovanovic, D.A. Jones, 2012: On the sensitivity of Australian temperature trends and variability to analysis methods and observation networks. CAWCR Technical Report No.050. Centre for Australian Weather and Climate Research, Melbourne, 54 pp. Feng, M., M.J. McPhaden, S.P. Xie, J. Hafner, 2013: La Nina forces unprecedented Leeuwin Current warming in 2011. Scientific Reports, 3, 1277. Fensham, R.J., R.J. Fairfax, D.P. Ward, 2009: Drought-induced tree death in savanna. Global Change Biology, 15(2), 380-387. Fidelman, P.I.J., A.M. Leitch, D.R. Nelson, 2013: Unpacking multilevel adaptation to climate change in the Great Barrier Reef, Australia. Global Environmental Change, 23(4), 800-812. Figueira, W.F., P. Biro, D.J. Booth, V.C. Valenzuela, 2009: Performance of tropical fish recruiting to temperate habitats: role of ambient temperature and implications of climate change. Marine Ecology-Progress Series, 384, 231-239. Figueira, W.F., D.J. Booth, 2010: Increasing ocean temperatures allow tropical fishes to survive overwinter in temperate waters. Global Change Biology, 16(2), 506-516. Finger, R., P. Lazzarotto, P. Calanca, 2010: Bio-economic assessment of climate change impacts on managed grassland production. Agricultural Systems, 103(9), 666-674. Finlay, K.J., A.L. Yen, J.P. Aurambout, G.A.C. Beattie, P. Barkley, J.E. Luck, 2009: Consequences for Australian biodiversity with establishment of the asiatic citrus psyllid, diaphorina citri, under present and future climates. Biodiversity, 10(2-3), 25-32. Fitzgerald, C., 2009: Implementing residential water use targets a Melbourne perspective. In: Efficient2009: 5th IWA Specialist Conference on Efficient Use and Management of Urban Water Supply, 25-28 October 2009, International Water Association, Sydney, 7 pp. Fitzgerald, G., R. Norton, M. Tausz, G. O Leary, S. Seneweera, S. Posch, M. Mollah, J. Brand, R. Armstrong, N. Mathers, J. Luck, W. Griffiths, P. Trebicki, 2010: Future effects of elevated CO2 on wheat production an overview of FACE research in Victoria, Australia. In: Food Security from Sustainable Agriculture. 15th Agronomy Conference, 15-18 November 2010, Lincoln, New Zealand [Dove, H., Culvenor, R.A. (eds.)], Australian Society of Agronomy, Gosford, NSW. Fitzharris, B.B., 2004: Possible impact of future climate change on seasonal snow of the Southern Alps of New Zealand. In: A Gaian World: Essays in Honour of Peter Holland [Fitzharris, G.K.a.B. (eds.)]. Department of Geography, University of Otago, Dunedin, pp. 231-241 Fitzharris, B.B., 2010: Climate change impacts on Dunedin. Dunedin City Council, Dunedin, 58 pp. Fitzpatrick, M.C., A.D. Gove, N.J. Sanders, R.R. Dunn, 2008: Climate change, plant migration, and range collapse in a global biodiversity hotspot: the Banksia (Proteaceae) of Western Australia. Global Change Biology, 14(6), 1337-1352. Fitzsimons, P., M. Bond, S. Webber, 2010: Creating a participatory adaptive capacity index for climate change adaptation - Report of engagement process in the South-West of Victoria. Department of Primary Industries, State Government of Victoria, Tatura, Vic, 41 pp. Fletcher, C.S., B.M. Taylor, A.N. Rambaldi, B.P. Harman, S. Heyenga, K.R. Ganegodage, F. Lipkin, R.R.J. McAllister, 2013: Costs and coasts: an empirical assessment of physical and institutional climate adaptation pathways. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, 53 pp. FOA, MPI, 2012: New Zealand Plantation Forestry Industry: Facts and Figures 2011/2012. New Zealand Forest Owners Association Inc. and Ministry for Primary Industries, Wellington, 46 pp. Fouquet, A., G.F. Ficetola, A. Haigh, N. Gemmell, 2010: Using ecological niche modelling to infer past, present and future environmental suitability for Leiopelma hochstetteri, an endangered New Zealand native frog. Biological Conservation, 143(6), 1375-1384. Frame, B., M. Brignall-Theyer, R. Taylor, K. Delaney, 2007: Work in progress: four future scenarios for New Zealand. Manaaki Whenua Press, Lincoln, New Zealand, 112 pp. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Frederiksen, J.S., C.S. Frederiksen, 2007: Interdecadal changes in southern hemisphere winter storm track modes. Tellus Series a-Dynamic Meteorology and Oceanography, 59(5), 599-617. Frederiksen, C.S., J.S. Frederiksen, J.M. Sisson, S.L. Osbrough, 2011: Changes and projections in Australian winter rainfall and circulation: Anthropogenic forcing and internal variablilty. Int. J. Climate Change Impacts and Responses, 2, 143-162. Freeman, C., C. Cheyne, 2008: Coasts for Sale: gentrification in New Zealand. Planning Theory and Practice, 9(1), 33-56. Fritze, J.G., G.A. Blashki, S. Burke, J. Wiseman, 2008: Hope, despair and transformation: climate change and the promotion of mental health and wellbeing. International Journal of Mental Health Systems, 2(1), 13. Fuentes, M.M.P.B., M. Hamann, V. Lukoschek, 2009: Marine Reptiles. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 18 pp. Fulton, E.A., 2011: Interesting times: winners, losers, and system shifts under climate change around Australia. Ices Journal of Marine Science, 68(6), 1329-1342. Fünfgeld, H., B. Webb, D. McEvoy, 2012: The Signi cance of Adaptation Framing in Local and Regional Climate Change Adaptation Initiatives in Australia. In: Resilient Cities 2. Cities and Adaptation to Climate Change Proceedings of the Global Forum 2011 [Otto-Zimmermann, K. (eds.)]. Springer, Dordrecht, the Netherlands, pp. 283-294. Funk, J., S. Kerr, 2007: Restoring Forests through Carbon Farming on M ori Land in New Zealand/Aotearoa. Mountain Research and Development, 27(3), 202-205. Gallagher, R.V., D. Englert Duursma, J. O'Donnell, P.D. Wilson, P.O. Downey, L. Hughes, M.R. Leishman, 2012a: The grass may not always be greener: projected reductions in climatic suitability for exotic grasses under future climates in Australia. Biological Invasions, advance online, DOI 10.1007/s10530-012-0342-6. Gallagher, R.V., L. Hughes, M.L. Leishman, 2012b: Species loss and gain in communities under future climate change: consequences for functional diversity Ecography, 36, 531-540. Gallant, A.J.E., K.J. Hennessy, J. Risbey, 2007: Trends in rainfall indices for six Australian regions: 1910-2005. Australian Meteorological Magazine, 56, 223-239. Gallant, A.J.E., D.J. Karoly, 2010: A Combined Climate Extremes Index for the Australian Region. Journal of Climate, 23, 6153-6165. Gallant, A.J.E., A.S. Kiem, D.C. Verdon-Kidd, R.C. Stone, D.J. Karoly, 2012: Understanding hydroclimate processes in the Murray-Darling Basin for natural resources management. Hydrology and Earth System Sciences, 16(7), 2049-2068. Gardiner, L., D. Firestone, G. Waibl, N. Mistal, K. Van Reenan, D.Hynes, J. Byfield, S. Oldfield, S. Allan, B. Kouvelis, J. Smart, A. Tait, A. Clark, 2009: Climate Change Effects on the Land Transport Network Volume One: Literature Review and Gap Analysis. Report No 378. NZ Transport Agency, Wellington, New Zealand, 226 pp. Gardner, J., A.-M. Dowd, C. Mason, P. Ashworth, 2009a: A framework for stakeholder engagement on climate adaptation. CSIRO Climate Adaptation Flagship Working paper No. 3. CSIRO, Kenmore, Qld, 32 pp. Gardner, J.L., R. Heinsohn, L. Joseph, 2009b: Shifting latitudinal clines in avian body size correlate with global warming in Australian passerines. Proceedings of the Royal Society B-Biological Sciences, 276(1674), 3845- 3852. Gardner, J., R. Parsons, G. Paxton, 2010: Adaptation benchmarking survey: initial report. CSIRO Climate Adaptation Flagship Working paper No. 4. CSIRO, Canberra, 58 pp. Garnaut, R., 2008: The Garnaut Climate Change Review: Final Report. Cambridge University Press, Melbourne, 634 pp. Garnett, S., D. Franklin, G. Ehmke, J. VanDerWal, L. Hodgson, C. Pavey, A. Reside, J. Welbergen, S. Butchart, G. Perkins, S. Williams, 2013: Climate change adaptation strategies for Australian birds. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 109 pp. Gaydon, R.S., H.G. Beecher, R. Reinke, S. Crimp, S.M. Howden, 2010: Rice. In: Adapting Agriculture to Climate Change. Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 67-83. GBRMPA, 2007: Great Barrier Reef Climate Change Action Plan 2007-2012. Great Barrier Reef Marine Park Authority, Commonwealth of Australia, Townsville, Qld, 14 pp. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 GBRMPA, 2009a: Great Barrier Reef Outlook Report 2009. Great Barrier Reef Marine Park Authority, Townsville, Qld, 212 pp. GBRMPA, 2009b: Great Barrier Reef Tourism Climate Change Action Strategy 2009-2012. Great Barrier Reef Marine Park Authority, Townsville, Qld, 56 pp. GBRMPA, 2011: Great Barrier Reef Report Card 2011: Reef Water Quality Protection Plan. Queensland Government, Brisbane, Qld, 6 pp. Gerard, P.J., J.R.F. Barringer, J.G. Charles, S.V. Fowler, J.M. Kean, C.B. Phillips, A.B. Tait, G.P. Walker, 2012: Potential effects of climate change on biological control systems: case studies from New Zealand. BioControl, advance online, DOI: 10.1007/s10526-012-9480-0. GETF, 2011: Roadmap to renewable and low emission energy in remote communities. Report by the Green Energy Task Force. Territory Climate Change, Northern Territory Government, Darwin, 65 pp. Gibbs, M., T. Hill, 2012: Coastal Climate Change Risk - Legal and Policy Responses in Australia. Department of Climate Change and Energy Efficiency, Canberra, 89 pp. Gibson, L., A. McNeill, P. de Tores, A. Wayne, C. Yates, 2010: Will future climate change threaten a range restricted endemic species, the quokka (Setonix brachyurus), in south west Australia? Biological Conservation, 143(11), 2453-2461. Gifford, R., L. Scannell, C. Kormos, L. Smolova, A. Biel, S. Boncu, V. Corral, H. Güntherf, K. Hanyu, D. Hine, F.G. Kaiser, K. Korpela, L.M. Lima, A.G. Mertig, R.G. Mira, G. Moser, P. Passafaro, J.Q. Pinheiro, S. Saini, T. Sako, E. Sautkina, Y. Savina, P. Schmuck, W. Schultz, K. Sobeck, E.-L. Sundblad, D. Uzzell, 2009: Temporal pessimism and spatial optimism in environmental assessments: An 18-nation study. Journal of Environmental Psychology, 29(1), 1-12. Gifford, R., 2011: The Dragons of Inaction. Psychological Barriers That Limit Climate Change Mitigation and Adaptation. American Psychologist, 66(4), 290-302. Gillanders, B.M., T.S. Elsdon, I.A. Halliday, G.P. Jenkins, J.B. Robins, F.J. Valesini, 2011: Potential effects of climate change on Australian estuaries and fish utilising estuaries: a review. Marine and Freshwater Research, 62(9), 1115-1131. Giltrap, D., A.-G. Ausseil, J. Ekanayake, S.M. Pawson, P. Hall, P. Newsome, J. Dymond, 2009: Chapter 2: Environmental impacts of large-scale forestry for bioenergy. In: Bioenergy options for New Zealand. Analysis of large-scale bioenergy from forestry [Hall, P., Jack, M. (eds.)]. Scion, Rorotua, NZ, pp. 71-121. Glavovic, B., W. Saunders, J. Becker, 2010: Land-use planning for natural hazards in New Zealand: the setting, barriers, burning issues and priority actions. Natural Hazards, 54(3), 679-706. Gorddard, R., R. Wise, K. Alexander, A. Langston, A. Leitch, M. Dunlop, A. Ryan, J. Langridge, 2012: Striking the balance: coastal development and ecosystem values. Department of Climate Change and Energy Efficiency, Canberra, 114 pp. Gordon, D.P., J. Beaumont, A. MacDiarmid, D.A. Robertson, S.T. Ahyong, 2010: Marine Biodiversity of Aotearoa New Zealand. Plos One, 5(8), e10905. Gorman-Murray, A., 2008: Before and after Climate Change: The Snow Country in Australian Imaginaries. M/C Journal, 11(5), online. Gorman-Murray, A., 2010: An Australian Feeling for Snow: Towards Understanding Cultural and Emotional Dimensions of Climate Change. Cultural Studies Review, 16(1), 60-81. Gosling, S.N., G.R. McGregor, J.A. Lowe, 2009: Climate change and heat-related mortality in six cities. Part 2: climate model evaluation and projected impacts from changes in the mean and variability of temperature with climate change. International Journal of Biometeorology, 53(1), 31-51. Gosling, E., K.J.H. Williams, 2010: Connectedness to nature, place attachment and conservation behaviour: Testing connectedness theory among farmers. Journal of Environmental Psychology, 30(3), 298-304. Goss, J., B. Lindquist, 2000: Placing movers: An overview of the Asia-Oacific migration system. Contemporary Pacific, 12(2), 385-414. Grace, W.J., V.O. Sadras, P.T. Hayman, 2009: Modelling heatwaves in viticultural regions of southeastern Australia. Australian Meteorological and Oceanographic Journal, 58(4), 249-262. Grafton, R.Q., K. Hussey, 2007: Buying back the living Murray: at what price? Australasian Journal of Environmental Management, 14(2), 74-81. Graham, P., L. Reedman, F. Poldy, 2008: Modelling of the future of transport fuels in Australia. IR 1046. CSIRO, Canberra, 112 pp. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Gray, M., J.C. Altman, N. Halasz, 2005a: The economic value of wild resources to the Indigenous community of the Wallis Lake catchment. CAEPR Discussion Paper No 272/2005. Centre for Aboriginal Economic Policy Research, The Australian National University, Canberra, 40 pp. Gray, W., R. Ibbitt, R. Turner, M. Duncan, M. Hollis, 2005b: A Methodology to assess the impacts of climate change on flood risk in New Zealand. NIWA Client Report CHC2005-060. Ministry for the Environment, Wellington, 40 pp. Green, A.E., M.R. Power, D.M. Jang, 2008a: Trans-Tasman migration: New Zealanders explanations for their move. New Zealand Geographer, 64(1), 34-45. Green, K., J.A. Stein, M.M. Driessen, 2008b: The projected distributions of Mastacomys fuscus and Rattus lutreolus in south-eastern Australia under a scenario of climate change: potential for increased competition? Wildlife Research, 35(2), 113-119. Green, D., S. Jackson, J. Morrison, 2009: Risks from climate change to Indigenous communities in the tropical north of Australia. Department of Climate Change and Energy Efficiency, Canberra, 185 pp. Green, B.S., C. Gardner, A. Linnane, P.J. Hawthorne, 2010a: The Good, the Bad and the Recovery in an Assisted Migration. Plos One, 5(11), e14160. Green, D., L. Alexander, K. McLnnes, J. Church, N. Nicholls, N. White, 2010b: An assessment of climate change impacts and adaptation for the Torres Strait Islands, Australia. Climatic Change, 102(3-4), 405-433. Green, D., J. Billy, A. Tapim, 2010c: Indigenous Australians' knowledge of weather and climate. Climatic Change, 100(2), 337-354. Griffiths, G.M., L.E. Chambers, M.R. Haylock, M.J. Manton, N. Nicholls, H.J. Baek, Y. Choi, P.M. Della-Marta, A. Gosai, N. Iga, R. Lata, V. Laurent, L. Maitrepierre, H. Nakamigawa, N. Ouprasitwong, D. Solofa, L. Tahani, D.T. Thuy, L. Tibig, B. Trewin, K. Vediapan, P. Zhai, 2005: Change in mean temperature as a predictor of extreme temperature change in the Asia-Pacific region. International Journal of Climatology, 25(10), 1301- 1330. Griffiths, G.M., 2007: Changes in New Zealand daily rainfall extremes 1930 2004. Weather and Climate, 27, 3-44. Guillaume, J., G.M. Li, M.F. Hutchinson, K. Proust, S. Dovers, 2010: Integrated Assessment of Climate Change Impacts on Urban Settlements (IACCIUS) Project: A differential vulnerability assessment of climate change impacts on Darwin. Fenner School of Environment and Society, The Australian National University, Canberra. Gunasekera, D., Y. Kim, C. Tulloh, M. Ford, 2007: Climate Change: impacts on Australian agriculture. Australian Commodities, 14, 657-676. Gunasekera, D., M. Ford, E. Heyhoe, A. Gurney, H. Ahammad, S.J. Phipps, I.N. Harman, J.J. Finnigan, M. Brede, 2008: Global integrated assessment model: a new analytical tool for assessing climate change risks and policies. Australian Commodities, 15(1), 195-216. Gurran, N., C. Squires, E.J. Blakely, 2006: Meeting the sea change challenge: best practice models of local & regional planning for sea change communities. Report No. 2 for the National Sea Change Taskforce. The University of Sydney Planning Research Centre, Sydney, 130 pp. Gurran, N., 2008: The Turning Tide: Amenity Migration in Coastal Australia. International Planning Studies, 13(4), 391-414. Gurran, N., E. Hamin, B. Norman, 2008: Planning for climate change: Leading Practice Principles and Models for Sea Change Communities in Coastal Australia. Prepared for the National Sea Change Taskforce. The University of Sydney, Faculty of Architecture, Sydney, 66 pp. Hagger, V., D. Fisher, S. Schmidt, S. Blomberg, 2013: Assessing the vulnerability of an assemblage of subtropical rainforest vertebrate species to cliamte change in south-east Queensland. Austral Ecology, 38, 465-475. Hall, A., G.V. Jones, 2009: Effect of potential atmospheric warming on temperature-based indices describing Australian winegrape growing conditions. Australian Journal of Grape and Wine Research, 15(2), 97-119. Hallegraeff, G.M., 2010: Ocean climate change, phytoplankton community responses, and harmful algal blooms: a formidable predictive challenge. Journal of Phycology, 46(2), 220-235. Hamin, E.M., N. Gurran, 2009: Urban form and climate change: Balancing adaptation and mitigation in the US and Australia. Habitat International, 33(3), 238-245. Handmer, J., K. Haynes, 2008: Community Bushfire Safety. CSIRO Publishing, Collingwood, Vic, 208 pp. Hanigan, I.C., C.D. Butler, P.N. Kokic, M.F. Hutchinson, 2012: Suicide and drought in New South Wales, Australia, 1970-2007. Proceedings of the National Academy of Sciences, 109(35), 13950-13955. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hanna, E.G., T. Kjellstrom, C. Bennett, K. Dear, 2011: Climate change and rising heat: population health implications for working people in Australia. Asia Pacific Journal of Public Health, 23(Supplement 2), S14- 26. Hannah, J., 2004: An updated analysis of long-term sea level change in New Zealand. Geophysical Research Letters, 31(3), L03307. Hannah, J., R. Bell, 2012: Regional sea level trends in New Zealand. Journal of Geophysical Research-Oceans, 117, C01004. Hansen, A., P. Bi, M. Nitschke, P. Ryan, D. Pisaniello, G. Tucker, 2008: The effect of heat waves on mental health in a temperate Australian city. Environmental Health Perspectives, 116(10), 1369-1375. Hanslow, K., D. Gunasekera, B.R. Cullen, D. Newth, 2013: Economic impacts of climate change on the Australian dairy sector. Australian Journal of Agricultural and Resource Economics, advance online, DOI: 10.1111/1467- 8489.12021. Harley, D., P. Bi, G. Hall, A. Swaminathan, S.L. Tong, C. Williams, 2011: Climate Change and Infectious Diseases in Australia: Future Prospects, Adaptation Options, and Research Priorities. Asia-Pacific Journal of Public Health, 23(2, Supplement), 54-66. Harman, I.N., M. Ford, G. Jakeman, S.J. Phipps, M. Brede, J.J. Finnigan, D. Gunasekera, H. Ahammad, 2008: Assessment of future global scenarios for the Garnaut Climate Change Review: an application of the GIAM framework. CSIRO, Canberra, 64 pp. Harmsworth, G., 2003: M ori perspectives on Kyoto policy: Interim Results- Reducing Greenhouse Gas Emissions from the Terrestrial Biosphere (C09X0212). Landcare Research NZ Ltd, Palmerston North, 33 pp. Harmsworth, G., B. Raynor, 2005: Cultural Consideration in Landslide Risk Perception. In: Landslide Hazard and Risk [Glade, T., Anderson, M.G., Crozier, M.J. (eds.)]. John Wiley & Sons, Ltd, London, pp. 219-249. Harmsworth, G., M. Tahi, C.K. Insley, 2010: Climate change business opportunities for M ori land and M ori organisations: Sustainable Land Management Mitigation and Adaptation to Climate Change (SLMACC). Manaaki Whenua Press, Palmerston North, 59 pp. Harper, B., T. Hardy, L. Mason, R. Fryar, 2009: Developments in storm tide modelling and risk assessment in the Australian region. Natural Hazards, 51(1), 225-238. Harris, R.J., G. Barker, 2007: Relative risk of invasive ants (Hymenoptera: Formicidae) establishing in New Zealand. New Zealand Journal of Zoology, 34(3), 161-178. Harris, J.B.C., D.A. Fordham, P.A. Mooney, L.P. Pedler, M.B. Araujo, D.C. Paton, M.G. Stead, M.J. Watts, H.R. Akcakaya, B.W. Brook, 2012: Managing the long-term persistence of a rare cockatoo under climate change. Journal of Applied Ecology, 49(4), 785-794. Harsch, M.A., R. Buxton, R.P. Duncan, P.E. Hulme, P. Wardle, J. Wilmshurst, 2012: Causes of tree line stability: stem growth, recruitment and mortality rates over 15 years at New Zealand Nothofagus tree lines. Journal of Biogeography, 39(11), 2061-2071. Hartog, J.R., A.J. Hobday, R. Matear, M. Feng, 2011: Habitat overlap between southern bluefin tuna and yellowfin tuna in the east coast longline fishery - implications for present and future spatial management. Deep-Sea Research Part II - Topical Studies in Oceanography, 58(5), 746-752. Hassim, M.E.E., K.J.E. Walsh, 2008: Tropical cyclone trends in the Australian region. Geochemistry Geophysics Geosystems, 9(7), Q07V07. Hasson, A.E.A., G.A. Mills, B. Timbal, K. Walsh, 2009: Assessing the impact of climate change on extreme fire weather events over southeastern Australia. Climate Research, 39(2), 159-172. Hayward, B., 2008a: 'Nowhere Far From the Sea': political challenges of coastal adaptation to climate change in New Zealand. Political Science, 60(1), 47-59. Hayward, B., 2008b: Let s Talk about the Weather: Decentering Democratic Debate about Climate Change. Hypathia, 23(3), 79-98. Hayward, J.A., P.W. Graham, P.K. Campbell, 2011: Projections of the future costs of electricity generation technologies. EP104982. CSIRO, Canberra, 79 pp. HCCREMS, 2010: Potential Impacts of Climate Change on the Hunter, Central and Lower North Coast of NSW. Hunter Councils, Thornton, NSW, 86 pp. HCCREMS, 2012: Decision Support for Coastal Adaptation: The Handbook. Hunter Councils, Thornton, NSW, 178 pp. HDC, 2012: Long Term Plan 2012/22. Hastings District Council, Hastings, 80 pp. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Heckbert, S., J. Russell-Smith, A. Reeson, J. Davies, G. James, C. Meyer, 2012: Spatially explicit benefit cost analysis of fire management for greenhouse gas abatement. Austral Ecology, 37(6), 724-732. Hellmann, J.J., J.E. Byers, B.G. Bierwagen, J.S. Dukes, 2008: Five Potential Consequences of Climate Change for Invasive Species. Conservation Biology, 22(3), 534-543. Hendon, H.H., D.W.J. Thompson, M.C. Wheeler, 2007: Australian rainfall and surface temperature variations associated with the Southern Hemisphere annular mode. Journal of Climate, 20(11), 2452-2467. Hendrikx, J., E. Hreinsson, 2012: The potential impact of climate change on seasonal snow in New Zealand: part II industry vulnerability and future snowmaking potential. Theoretical and Applied Climatology, 110(4), 619- 630. Hendrikx, J., E. Hreinsson, M. Clark, B. Mullan, 2012: The potential impact of climate change on seasonal snow in New Zealand; Part I - An analysis using 12 GCMs. Theoretical and Applied Climatology, 110(4), 607-618. Hendrikx, J., C. Zammit, E. Hreinsson, S. Becken, 2013: A comparative assessment of the potential impact of climate change on the ski industry in New Zealand and Australia. Climatic Change, 119(3-4), 965-978. Hennessy, K.J., B. Fitzharris, B. Bates, N. Harvey, M. Howden, L. Hughes, J. Salinger, R. Warrick, 2007: Australia and New Zealand. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Parry, M., Canziani, O., Palutikof, J., van der Linden, P., Hanson, C. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 507-540. Hennessy, K.J., R. Fawcett, D.G.C. Kirono, F.S. Mpelasoka, D. Jones, J.M. Bathols, P.H. Whetton, M. Stafford Smith, M. Howden, C.D. Mitchell, N. Plummer, 2008a: An assessment of the impact of climate change on the nature and frequency of exceptional climatic events. CSIRO and Bureau of Meteorology, Melbourne and Canberra, 33 pp. Hennessy, K.J., P.H. Whetton, K. Walsh, I.N. Smith, J.M. Bathols, M. Hutchinson, J. Sharples, 2008b: Climate change effects on snow conditions in mainland Australia and adaptation at ski resorts through snowmaking. Climate Research, 35(3), 255-270. Hertzler, G., 2007: Adapting to climate change and managing climate risks by using real options. Australian Journal of Agricultural Research, 58(10), 985-992. Hine, D.W., W.J. Phillips, J.P. Reser, R.W. Cooksey, A.D.G. Marks, P.D. Nunn, S.E. Watt, M.C. Ellul, 2013: Enhancing climate change communication: strategies for profiling and targeting Australian interpretive communities. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 95 pp. Hitchings, T.R., 2009: Leptophlebiidae (Ephemeroptera) of the alpine region of the Southern Alps, New Zealand. Aquatic Insects, 31, 595-601. Hobday, A.J., E.S. Poloczanska, R. Matear (eds.), 2008: Implications of Climate Change for Australian Fisheries and Aquaculture: A preliminary assessment. Department of Climate Change, Canberra, 86 pp. Hobday, A.J., 2010: Ensemble analysis of the future distribution of large pelagic fishes off Australia. Progress in Oceanography, 86(1-2), 291-301. Hobday, A.J., E.S. Poloczanska, 2010: Fisheries and Aquaculture. In: Adapting Agriculture to Climate Change: Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 205-228. Hobday, A.J., 2011: Sliding baselines and shuffling species: implications of climate change for marine conservation. Marine Ecology, 32(3), 392-403. Hobson, K., S. Niemeyer, 2011: Public responses to climate change: The role of deliberation in building capacity for adaptive action. Global Environmental Change, 21(3), 957-971. Hodgkinson, J.H., A. Littleboy, M. Howden, K. Moffat, B. Loechel, 2010a: Climate adaptation in the Australian mining and exploration industries. CSIRO Climate Adaptation Flagship Working paper No. 5. CSIRO, Canberra, 32 pp. Hodgkinson, J.H., B. Loechel, K. Moffat, M. Howden, A. Littleboy, S. Crimp, 2010b: Climate risks to the Australian mining industry - a preliminary review of vunerabilities. In: Sustainable Mining 2010 Conference: 17-19 August 2010, Kalgoorlie, Western Australia. The Australasian Institute of Mining and Metallurgy, Carlton South, Vic, pp. 341-350. Hoegh-Guldberg, O., P.J. Mumby, A.J. Hooten, R.S. Steneck, P. Greenfield, E. Gomez, C.D. Harvell, P.F. Sale, A.J. Edwards, K. Caldeira, N. Knowlton, C.M. Eakin, R. Iglesias-Prieto, N. Muthiga, R.H. Bradbury, A. Dubi, M.E. Hatziolos, 2007: Coral reefs under rapid climate change and ocean acidification. Science, 318(5857), 1737-1742. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hoegh-Guldberg, O., 2012: The adaptation of coral reefs to climate change: is the Red Queen being outpaced? Scientia Marina, 76(2), 403-408. Hoelzle, M., T. Chinn, D. Stumm, F. Paul, M. Zemp, W. Haeberli, 2007: The application of glacier inventory data for estimating past climate change effects on mountain glaciers: A comparison between the European Alps and the Southern Alps of New Zealand. Global and Planetary Change, 56(1 2), 69-82. Hofmeester, C., B. Bishop, L. Stocker, G. Syme, 2012: Social cultural influences on current and future coastal governance. Futures, 44(8), 719-729. Hogan, A., H.L. Berry, S.P. Ng, A. Bode, 2011a: Decisions made by farmers related to climate change. Publication No. 10/28, project No. PRJ-004546. Rural Industries Research and Development Corporation, Barton, ACT, 64 pp. Hogan, A., A. Bode, H. Berry, 2011b: Farmer Health and Adaptive Capacity in the Face of Climate Change and Variability. Part 2: Contexts, Personal Attributes and Behaviors. International Journal of Environmental Research and Public Health, 8(10), 4055-4068. Holbrook, N.J., J. Davidson, M. Feng, A.J. Hobday, J.M. Lough, S. McGregor, J.S. Risbey, 2009: El Nino- Southern Oscillation. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 25 pp. Hope, P., W. Drosdowsky, N. Nicholls, 2006: Shifts in the synoptic systems influencing southwest Western Australia. Climate Dynamics, 26(7-8), 751-764. Hope, P., B. Timbal, R. Fawcett, 2010: Associations between rainfall variability in the southwest and southeast of Australia and their evolution through time. International Journal of Climatology, 30(9), 1360-1371. Hopkins, D., J.E.S. Higham, S. Becken, 2012: Climate change in a regional context: relative vulnerability in the Australasian skier market. Regional Environmental Change, advance online, DOI 10.1007/s10113-012-0352-z. Hovenden, M.J., A.L. Williams, 2010: The impacts of rising CO2 concentrations on Australian terrestrial species and ecosystems. Austral Ecology, 35(6), 665-684. Howard, W.R., M. Nash, K. Anthony, K. Schmutter, H. Bostock, D. Bromhead, M. Byrne, K. Currie, G. Diaz Pulido, S. Eggins, M. Ellwood, B. Eyre, R. Haese, G. Hallegraeff, K. Hill, C. Hurd, C. Law, A. Lenton, R. Matear, B. McNeil, M. McCulloch, M.N. Müller, P. Munday, B. Opdyke, J.M. Pandolfi, R. Richards, D. Roberts, B.D. Russell, A.M. Smith, B. Tilbrook, A. Waite, J. Williamson, 2012: Ocean acidification. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2012 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 38 pp. Howden, S.M., S.J. Crimp, 2008: Drought and High Temperature Extremes: Effects on Water and Electricity Demands. In: Transitions: Pathways Towards Sustainable Urban Development in Australia [Newton, P.W. (eds.)]. CSIRO, Collingwood, Vic, pp. 227-244. Howden, S.M., S.J. Crimp, C.J. Stokes, 2008: Climate change and Australian livestock systems: Impacts, research and policy issues. Australian Journal of Experimental Agriculture, 48(7), 780-788. Howden, S.M., S.J. Crimp, R. Nelson, 2010: Australian agriculture in a climate of change. In: Managing Climate Change: Papers from the GREENHOUSE 2009 Conference [Jubb, I., Holper, P., Cai, W. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 101-111. Howden, S.M., C.J. Stokes, 2010: Introduction. In: Adapting Agriculture to Climate Change: Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 1-11. Howden-Chapman, P., 2010: Climate change & human health: Impact and adaptation issues for New Zealand. In: Climate Change Adaptation in New Zealand: Future scenarios and some sectoral perspectives [Nottage, R., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 112-121. Howe, C., R.N. Jones, S. Maheepala, B. Rhodes, 2005: Melbourne Water Climate Change Study: Implications of Potential Climate Change for Melbourne s Water Resources. Report CMIT-2005-104. CSIRO and Melbourne Water, Melbourne, 26 pp. Howells, E.J., R. Berkelmans, M.J.H. van Oppen, B.L. Willis, L.K. Bay, 2013: Historical thermal regimes define limits to coral acclimatization. Ecology, 94(5), 1078-1088. HRC, 2012: Proposed One Plan. Chapter 10: Natural Hazards. Horizons Regional Council, Palmerston North, 8 pp. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Huang, C., A.G. Barnett, X. Wang, S. Tong, 2012: The impact of temperature on years of life lost in Brisbane, Australia. Nature Clim. Change, 2(4), 265-270. Hughes, L., 2003: Climate change and Australia: trends, projections and impacts. Austral Ecology, 28, 423-443. Hughes, T.P., M.J. Rodrigues, D.R. Bellwood, D. Ceccarelli, O. Hoegh-Guldberg, L. McCook, N. Moltschaniwskyj, M.S. Pratchett, R.S. Steneck, B. Willis, 2007: Phase Shifts, Herbivory, and the Resilience of Coral Reefs to Climate Change. Current Biology, 17(4), 360-365. Hughes, R., D. Mercer, 2009: Planning to Reduce Risk: The Wildfire Management Overlay in Victoria, Australia. Geographical Research, 47(2), 124-141. Hughes, L., R. Hobbs, A. Hopkins, J. McDonald, M. Stafford-Smith, W. Steffen, S.E. Williams, 2010: National Climate Change Adaptation Research Plan: Terrestrial Biodiversity. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 70 pp. Hughes, J.D., K.C. Petrone, R.P. Silberstein, 2012: Drought, groundwater storage and streamflow decline in southwestern Australia. Geophysical Research Letters, 39, L03408. Hugo, G., 2010: Climate change-induced mobility and the existing migration regime in Asia and the Pacific. In: Climate Change and Displacement [McAdam, J. (eds.)]. Hart Publishing, Oxford, pp. 9-36. Hugo, G., 2012: Population Distribution, Migration and Climate Change in Australia: An Exploration. ACCARNSI Discussion paper - Node 2. Urban management, transport and social inclusion. National Adaptation Research Facility (NCCARF), Gold Coast, Qld, 101 pp. Hunt, J., J. Altman, K. May, 2009: Social benefits of Aboriginal engagement in natural resource management. CAEPR Working Paper 60. Centre for Aboriginal Economic Policy Research, The Australian National University, Canberra, 88 pp. Hunter, E., 2009: 'Radical hope' and rain: Climate change and the mental health of Indigenous residents of northern Australia. Australasian Psychiatry, 17(6), 445-452. Hurlimann, A., S. Dolnicar, 2011: Voluntary relocation - An exploration of Australian attitudes in the context of drought, recycled and desalinated water. Global Environmental Change, 21(3), 1084-1094. Hussey, K., S. Dovers (eds.), 2007: Managing water for Australia: the social and institutional challenges. CSIRO Publishing, Collingwood, Vic, 157 pp. IAA, 2011a: Institute of Actuaries of Australia, submission No. 43 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. Productivity Commission, Canberra (online at www.pc.gov.au/projects/inquiry/climate-change-adaptation/submissions), 60 pp. IAA, 2011b: Institute of Actuaries of Australia, submission to the Garnaut Climate Change Review - Update 2011. Garnaut Climate Change Review, Canberra, 28 pp. IAG, 2011: Insurance Australia Group, submission No. 39 (2011) to Productivity Commission Issues Paper: Barriers to Effective Climate Change Adaptation. Productivity Commission, Canberra (online at www.pc.gov.au/projects/inquiry/climate-change-adaptation/submissions), 24 pp. ICA, 2012: Historical Disaster Statistics (online database, accessed 31 August 2013, at http://www.insurancecouncil.com.au/industry-statistics-data/disaster-statistics/historical-disaster-statistics). Insurance Council of Australia, Sydney, NSW. ICNZ, 2013: The Cost of Disaster events (online database, accessed 31 August 2013, at www.icnz.org.nz/natural- disaster/historic-events). Insurance Council of New Zealand, Wellington. Insley, C.K., 2007: M ori impacts from emissions trading scheme. Interim high level findings. 37 Degrees South Aotearoa, Gisborne, New Zealand, 22 pp. Insley, C.K., R. Meade, 2008: M ori Impacts from the Emissions Trading Scheme: Detailed Analysis and Conclusions. Ministry for the Environment, Wellington, New Zealand, 57 pp. Insley, C.K., 2010: Survey of M ori business: climate change M ori business opportunities. 37 Degrees South Aotearoa, Gisborne, New Zealand, 35 pp. IOCI, 2012: Indian Ocean Climate Initiative Stage 3: Summary for Policymakers. CSIRO and BOM, Melbourne, 43 pp. IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, UK, 582 pp. Irving, A.D., S.D. Connell, B.D. Russell, 2011: Restoring Coastal Plants to Improve Global Carbon Storage: Reaping What We Sow. Plos One, 6(3), e18311. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Irving, D., P. Whetton, A. Moise, 2012: Climate projections for Australia: a first glance at CMIP5. Australian Meteorological and Oceanographic Journal, 62(4), 211-225. Ishak, E.H., A. Rahman, S. Westra, A. Sharma, G. Kuczera, 2013: Evaluating the non-stationarity of Australian annual maximum flood. Journal of Hydrology, 494, 134 145. Jackson, A.C., M.G. Chapman, A.J. Underwood, 2008: Ecological interactions in the provision of habitat by urban development: whelks and engineering by oysters on artificial seawalls. Austral Ecology, 33(3), 307-316. Jakes, P.J., L. Kelly, E.R. Langer, 2010: An exploration of a fire-affected community undergoing change in New Zealand. Australian Journal of Emergency Management, 25(3), 48-53. Jakes, P.J., E.R. Langer, 2012: The adaptive capacity of New Zealand communities to wildfire. International Journal of Wildland Fire, 21(6), 764-772. Jakob, D., D. Karoly, A. Seed, 2011: Non-stationarity in daily and sub-daily intense rainfall - Part 2: Regional assessment for sites in south-east Australia. Natural Hazards and Earth System Sciences, 11(8), 2273-2284. Jellyman, D.J., D.J. Booker, E. Watene, 2009: Recruitment of Anguilla spp. glass eels in the Waikato River, New Zealand. Evidence of declining migrations? Journal of Fish Biology, 74(9), 2014-2033. Jenkins, G.P., S.D. Conron, A.K. Morison, 2010: Highly variable recruitment in an estuarine fish is determined by salinity stratification and freshwater flow: implications of a changing climate. Marine Ecology Progress Series, 417, 249-261. Jenkins, K.M., R.T. Kingsford, G.P. Closs, B.J. Wolfenden, C.D. Matthaei, S.E. Hay, 2011: Climate change and freshwater systems in Oceania: an assessment of vulnerability and adaptation opportunities. Pacific Conservation Biology, 17(3), 201-219. Johnson, C.R., S.C. Banks, N.S. Barrett, F. Cazassus, P.K. Dunstan, G.J. Edgar, S.D. Frusher, C. Gardner, M. Haddon, F. Helidoniotis, K.L. Hill, N.J. Holbrook, G.W. Hosie, P.R. Last, S.D. Ling, J. Melbourne-Thomas, K. Miller, G.T. Pecl, A.J. Richardson, K.R. Ridgway, S.R. Rintoul, D.A. Ritz, D.J. Ross, J.C. Sanderson, S.A. Shepherd, A. Slotvvinski, K.M. Swadling, N. Taw, 2011: Climate change cascades: Shifts in oceanography, species' ranges and subtidal marine community dynamics in eastern Tasmania. Journal of Experimental Marine Biology and Ecology, 400(1-2), 17-32. Johnston, F.H., D.M. Kavanagh, D.M. Bowman, R.K. Scott, 2002: Exposure to bushfire smoke and asthma: an ecological study. Medical Journal of Australia, 176, 535-538. Johnston, G., D. Burton, M. Baker-Jones, 2013: Climate Change Adaptation in the Boardroom: combined report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 85 pp. Jollands, N., M. Ruth, C. Bernier, N. Golubiewski, 2007: The climate's long-term impact on New Zealand infrastructure (CLINZI) project - A case study of Hamilton City, New Zealand. Journal of Environmental Management, 83(4), 460-477. Jones, R.N., F. Chiew, W. Boughton, L. Zhang, 2006: Estimating the sensitivity of mean annual runoff to climate change using selected hydrological models. Advances in Water Resources, 29(10), 1419-1429. Jones, D., W. Wang, R. Fawcett, 2009: High-quality spatial climate data-sets for Australia. Australian Meteorological and Oceanographic Journal, 58(4), 233-248. Jones, R., A. Wardell-Johnson, M. Gibberd, A. Pilgrim, G. Wardell-Johnson, J. Galbreath, S. Bizjak, D. Ward, K. Benjamin, J. Carlsen, 2010: The impact of climate change on the Margaret River Wine Region: developing adaptation and response strategies for the tourism industry. Sustainable Tourism Cooperative Research Centre, Gold Coast, Qld, 94 pp. Kamber, G., C. McDonald, G. Price, 2013: Drying out: Investigating the economic effects of drought in New Zealand. Reserve Bank of New Zealand Analytical Note series AN2013/02. Reserve Bank of New Zealand, Wellington, 31 pp. Kamman, C., L. Grünhage, U. Grüters, S. Janze, H.-J. Jäger, 2005: Response of aboveground grassland biomass and soil moisture to moderate long-term CO2 enrichment. Basic and Applied Ecology, 6, 351-365. Karoly, D.J., K. Braganza, 2005: Attribution of recent temperature changes in the Australian region. Journal of Climate, 18, 457-464. Karoly, D.J., S. Boulter, 2013: Afterword: floods, storms, fire and pestilence disaster risk in Australia during 2010/11. In: Natural Disasters and Adaptation to Climate Change [Boulter, S., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 412-418. Kawaguchi, S., A. Ishida, R. King, B. Raymond, N. Waller, A. Constable, S. Nicol, M. Wakita, A. Ishimatsu, 2013: Risk maps for Antarctic krill under projected Southern Ocean acidification. Nature Climate Change, advance online, DOI: 10.1038/nclimate1937. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kay, R., A. Travers, L. Dalton, 2014: The evolution of Coastal Vulnerability Assessments to Support Adaptive Decision Making in Australia: A Review. In: Climate Change and the Coastal Zone: Building Resilient Communities [Glavovic, B., Kay, R.C., Kelly, M., Travers, A. (eds.)]. CRC Press, London, www.crcpress.com/product/isbn/9780415464871. KCDC, 2012: Proposed Kapiti Coast District Plan, Volume 1. Notified 29 November 2012. Kapiti Coast District Council, Paraparaumu. Kearney, M.R., W.P. Porter, C. Williams, S. Ritchie, A.A. Hoffmann, 2009: Integrating biophysical models and evolutionary theory to predict climatic impacts on species ranges: the dengue mosquito Aedes aegypti in Australia. Functional Ecology, 23(3), 528-538. Kearney, M.R., N.J. Briscoe, D.J. Karoly, W.P. Porter, M. Norgate, P. Sunnucks, 2010a: Early emergence in a butterfly causally linked to anthropogenic warming. Biology Letters, 6(5), 674-677. Kearney, M.R., B.A. Wintle, W.P. Porter, 2010b: Correlative and mechanistic models of species distribution provide congruent forecasts under climate change. Conservation Letters, 3(3), 203-213. Keith, D.A., H.R. Akcakaya, W. Thuiller, G.F. Midgley, R.G. Pearson, S.J. Phillips, H.M. Regan, M.B. Araujo, T.G. Rebelo, 2008: Predicting extinction risks under climate change: coupling stochastic population models with dynamic bioclimatic habitat models. Biology Letters, 4(5), 560-563. Keith, D.A., S. Rodoreda, M. Bedward, 2010: Decadal change in wetland-woodland boundaries during the late 20th century reflects climatic trends. Global Change Biology, 16(8), 2300-2306. Kenderdine, S., 2010: Examining climate change: an Environment Court perspective. Resource Management Theory & Practice, 6, 35-92. Kennedy, D., L. Stocker, G. Burke, 2010: Australian local government action on climate change adaptation: some critical reflections to assist decision-making. Local Environment, 15(9/10), 805-816. Kenny, G., 2011: Adaptation in agriculture: lessons for resilience from eastern regions of New Zealand. Climatic Change, 106(3), 441-462. Kerr, S., W. Zhang, 2009: Allocation of New Zealand Units within Agriculture in the New Zealand Emissions Trading System. Working Paper 09-16. Motu Economic and Public Policy Research, Wellington, 82 pp. Khalaj, B., G. Lloyd, V. Sheppeard, K. Dear, 2010: The health impacts of heat waves in five regions of New South Wales, Australia: a case-only analysis. International Archives of Occupational and Environmental Health, 83(7), 833-842. Khan, S., 2012: Vulnerability assessments and their planning implications: a case study of the Hutt Valley, New Zealand. Natural Hazards, 64(2), 1587-1607. Kharin, V.V., F.W. Zwiers, X. Zhang, M. Wehner, 2013: Changes in temperature and precipitation extremes in the CMIP5 ensemble. Climatic Change, 119(2), 345-357. Kiem, A.S., S.W. Franks, G. Kuczera, 2003: Multi-decadal variability of flood risk. Geophysical Research Letters, 30(2), L015992. Kiem, A.S., D.C. Verdon-Kidd, S. Boulter, J. Palutikof, 2010: Learning from experience: Historical Case Studies and Climate Change Adaptation, Case studies of extreme events, Synthesis Report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 42 pp. Kiem, A.S., E.K. Austin, 2012: Limits and barriers to climate change adaptation for small inland communities affected by drought. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 46 pp. Kiem, A.S., E.K. Austin, 2013: Drought and the future of rural communities: Opportunities and challenges for climate change adaptation in regional Victoria, Australia. Global Environmental Change, advance online, DOI: 10.1016/j.gloenvcha.2013.06.003. King, D., W. Tawhai, A. Skipper, W. Iti, 2005: Anticipating local weather and climate outcomes using M ori environmental indicators. NIWA, Auckland, New Zealand, 18 pp. King, D., G. Penny, 2006: The 2nd M ori Climate Forum - Hongoeka Marae, Plimmerton (24 May 2006): Summary Report. National Institute of Water & Atmospheric Research Ltd, Auckland, New Zealand, 26 pp. King, D., J. Goff, A. Skipper, 2007: M ori environmental knowledge and natural hazards in Aotearoa New Zealand. Journal of the Royal Society of New Zealand, 37(2), 59-73. King, D., W. Iti, D. Hosking, 2008: Ground-truthing pre-event recovery planning issues with Ng ti Rongomai. NIWA, Auckland, New Zealand, 72 pp. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 King, D., G. Penny, C. Severne, 2010: The climate change matrix facing Maori society. In: Climate change adaptation in New Zealand: Future scenarios and some sectoral perspectives [Nottage, R., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 100-111. King, K.J., R.M. de Ligt, G.J. Cary, 2011: Fire and carbon dynamics under climate change in south-eastern Australia: insights from FullCAM and FIRESCAPE modelling. International Journal of Wildland Fire, 20(4), 563-577. King, D., W. Dalton, M. Home, M. Duncan, M.S. Srinivasan, J. Bind, C. Zammit, A. McKerchar, D. Ashford- Hosking, A. Skipper, 2012: M ori community adaptation to climate variability and change: Examining risks, vulnerability and adaptive strategies with Ng ti Huirapa at Arowhenua P , Te-umu-kaha (Temuka), New Zealand. Report AKL2011-015, prepared for Te R nanga o Arowhenua Society Incorporated and the New Zealand Climate Change Research Institute, Victoria University. NIWA, Auckland, 133 pp. Kingsford, R.T., J.E.M. Watson, C.J. Lundquist, O. Venter, L. Hughes, E.L. Johnston, J. Atherton, M. Gawel, D.A. Keith, B.G. Mackey, C. Morley, H.P. Possingham, B. Raynor, H.F. Recher, K.A. Wilson, 2009: Major Conservation Policy Issues for Biodiversity in Oceania. Conservation Biology, 23(4), 834-840. Kingsford, R.T. (ed.), 2011: Conservation Management of Rivers and Wetlands under Climate Change, Marine & Freshwater Research, Special Issue Volume 62 Number 3. CSIRO Publishing, Collingwood, Vic, 217-327 pp. Kingsford, R.T., J.E.M. Watson, 2011: Climate change in Oceania: A synthesis of biodiversity impacts and adaptations. Pacific Conservation Biology, 17, 270-284. Kinney, P., 2012: Winter mortality in a changing climate: will it go down? Bulletin épidémiologique hebdomadaire, 12-13, 148-151. Kirby, M., J. Connor, R. Bark, E. Qureshi, S. Keyworth, 2012: The economic impact of water reductions during the Millennium Drought in the Murray-Darling Basin. In: 56th AARES Annual Conference, 7-10 February 2012, Fremantle, Australia, Australian Agricultural & Resource Economics Society, Canberra, 25 pp. Kirby, M., F. Chiew, M. Mainuddin, B. Young, G. Podger, A. Close, 2013: Drought and climate change in the Murray-Darling Basin: a hydrological perspective. In: Drought in Arid and Semi-Arid Environments: A Multi- Disciplinary and Cross-Country Perspective [Schwabe, K., Albiac, J., Connor, J., Hassan, R., Meza-Gonzales, L. (eds.)]. Springer, Dordrecht, Netherlands, pp. 281-299. Kirono, D.G.C., D. Kent, 2010: Assessment of rainfall and potential evaporation from global climate modelsand its implications for Australian regional drought projection. International Journal of CLimatology, 31(9), 1295- 1308. Kirono, D.G.C., D.M. Kent, K.J. Hennessy, F.S. Mpelasoka, 2011: Characteristics of Australian droughts under enhanced greenhouse conditions: Results from 14 global climate models. Journal of Arid Environments, 75(6), 566-575. Kirschbaum, M.U.F., M.S. Watt, 2011: Use of a process-based model to describe spatial variation in Pinus radiata productivity in New Zealand. Forest Ecology and Management, 262, 1008-1019. Kirschbaum, M.U.F., N.W.H. Mason, M.S. Watt, A. Tait, A.E. Ausseil, D.J. Palmer, F.E. Carswell, 2011a: Productivity surfaces for Pinus radiata and a range of indigenous forest species under current climatic conditions. MAF Technical Paper No: 2011/45. Minstry of Agriculture and Forestry, Wellington, 129 pp. Kirschbaum, M.U.F., D. Whitehead, S.M. Dean, P.N. Beets, J.D. Shepherd, A.G.E. Ausseil, 2011b: Implications of albedo changes following afforestation on the benefits of forests as carbon sinks. Biogeosciences, 8(12), 3687- 3696. Kirschbaum, M.U.F., M.S. Watt, A. Tait, A.-G.E. Ausseil, 2012: Future productivity of Pinus radiata in New Zealand under expected climatic changes. Global Change Biology, 18(4), 1342-1356. Klamt, M., R. Thompson, J. Davis, 2011: Early response of the platypus to climate warming. Global Change Biology, 17(10), 3011-3018. Kouvelis, B., C. Scott, E. Rudkin, D. Cameron, M. Harkness, A. Renata, G. Williams, 2010: Impacts of Climate Change on Rural Water Infrastructure. Report to Ministry of Agriculture and Forestry. MWH New Zealand, Wellington, 122 pp. Kuczera, G., 1987: Prediction of water yield reductions following a bushfire in ash-mixed species eucalypt forest. Journal of Hydrology, 94(3-4), 215-236. Kulendran, N., L. Dwyer, 2010: Seasonal variation versus climate variation for Australian Tourism. Sustainable Tourism Cooperative Research Centre, Gold Coast, Qld, 27 pp. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kuleshov, Y., R. Fawcett, L. Qi, B. Trewin, D. Jones, J. McBride, H. Ramsay, 2010: Trends in tropical cyclones in the South Indian Ocean and the South Pacific Ocean. Journal of Geophysical Research-Atmospheres, 115, D01101. Kuruppu, N., J. Murta, P. Mukheibir, J. Chong, T. Brennan, 2013: Understanding the adaptive capacity of Australian small-to-medium enterprises to climate change and variability. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 133 pp. Kwan, D., H. Marsh, S. Delean, 2006: Factors influencing the sustainability of customary dugong hunting by a remote Indigenous community. Environmental Conservation, 33(2), 164-171. Lal, A., S. Hales, N. French, M.G. Baker, 2012: Seasonality in Human Zoonotic Enteric Diseases: A Systematic Review. Plos One, 7(4), e31883. Langton, M., M. Parsons, S. Leonard, K. Auty, D. Bell, C.P. Burgess, S. Edwards, R. Howitt, S. Jackson, V. McGrath, J. Morrison, 2012: National Climate Change Adaptation Research Plan for Indigenous Communities. NCCARF Publication 11/12. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 48 pp. Last, P.R., W.T. White, D.C. Gledhill, A.J. Hobday, R. Brown, G.J. Edgar, G. Pecl, 2011: Long-term shifts in abundance and distribution of a temperate fish fauna: a response to climate change and fishing practices. Global Ecology and Biogeography, 20(1), 58-72. Laurance, W.F., 2008: Global warming and amphibian extinctions in eastern Australia. Austral Ecology, 33(1), 1-9. Laurance, W.F., B. Dell, S.M. Turton, M.J. Lawes, L.B. Hutley, H. McCallum, P. Dale, M. Bird, G. Hardy, G. Prideaux, B. Gawne, C.R. McMahon, R. Yu, J.M. Hero, L. Schwarzkop, A. Krockenberger, M. Douglas, E. Silvester, M. Mahony, K. Vella, U. Saikia, C.H. Wahren, Z.H. Xu, B. Smith, C. Cocklin, 2011: The 10 Australian ecosystems most vulnerable to tipping points. Biological Conservation, 144(5), 1472-1480. Lawrence, J., S. Allan, 2009: A strategic framework and practical options for integrating flood risk management to reduce flood risk and the effects of climate change. Ministry for the Environment, Wellington, 79 pp. Lawrence, A., T. Baker, C.E. Lovelock, 2012: Optimising and managing coastal carbon: Comparative sequestration and mitigation opportunities across Australia s landscapes and land uses. FRDC Report 2011/084. Fisheries Research and Development Corporation, Deakin, ACT, 66 pp. Lawrence, J., A. Reisinger, B. Mullan, B. Jackson, 2013a: Exploring climate change uncertainties to support adaptive management of changing flood-risk. Environmental Science & Policy, 33, 133-142. Lawrence, J., F. Sullivan, A. Lash, G. Ide, C. Cameron, L. McGlinchey, 2013b: Adapting to changing climate risk by local government in New Zealand: institutional practice barriers and enablers. Local Environment, advance online, DOI: 10.1080/13549839.2013.839643. Leblanc, M., S. Tweed, A. Van Dijk, B. Timbal, 2012: A review of historical and future hydrological changes in the Murray-Darling Basin. Global and Planteray Change, 80-81, 226-246. Leitch, A.M., C.J. Robinson, 2012: Shifting sands: uncertainty and a local community response to sea level rise policy in Australia. In: Risk and Social Theory in Environmental Management: from uncertainty to local action [Measham, T., Lockie, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 117-131. Lennox, J., W. Proctor, S. Russell, 2011: Structuring stakeholder participation in New Zealand's water resource governance. Ecological Economics, 70(7), 1381-1394. Leonard, M., S. Westra, A. Phatak, M. Lambert, B. van den Hurk, K. McInnes, J. Risbey, S. Schuster, D. Jakob, M. Stafford-Smith, 2013a: Understanding the role of compound events in climate extremes. WIRES, advance online, DOI: 10.1002/wcc.252. Leonard, S., J. Mackenzie, F. Kofod, M. Parsons, M. Langton, P. Russ, L. Ormond-Parker, K. Smith, M. Smith, 2013b: Indigenous climate change adaptation in the Kimberley region of North-western Australiat: Learning from the past, adapting in the future: Identifying pathways to successful adaptation in Indigenous communities. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 131 pp. Leslie, L., M. Leplastrier, B. Buckley, 2008: Estimating future trends in severe hailstorms over the Sydney Basin: A climate modelling study. Atmospheric Research, 87(1), 37-51. Lett, R., R. Morden, C. McKay, T. Sheedy, M. Burns, D. Brown, 2009: Farm dam interception in the Campaspe Basin under climate change. In: H2009: 32nd Hydrology and Water Resources Symposium, 30 November-3 December 2009, Newcastle, Australia, Engineers Australia, Barton, ACT. pp 1194-1204. Leviston, Z., A. Leitch, M. Greenhill, R. Leonard, I. Walker, 2011: Australians views of climate change. CSIRO, Canberra, 21 pp. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Leviston, Z., I. Walker, S. Morwinski, 2012: Your opinion on climate change might not be as common as you think. Nature Clim. Change, 3, 334 337. Lewandowsky, S., 2011: Popular consensus: climate change is set to continue. Psychological Science, 22(4), 460- 463. LGNZ, 2008: Local Government New Zealand, submission on the proposed Coastal Policy Statement. Department of Conservation, Wellington, 40 pp. Lieffering, M., P.C.D. Newton, F.Y. Li, R. Vibart, 2012: Hill country sheep and beef: Impacts and adaptation to climate change In: Enhanced climate change impact and adaptation evaluation: A comprehensive analysis of New Zealand's land-based primary sectors [Clark, A., Nottage, R. (eds.)]. Ministry for Primary Industries, New Zealand, pp. 145-188. Lindenmayer, D.B., 2007: On borrowed time: Australia's environmental crisis and what we must do about it. Penguin Books in association with CSIRO Publishing, Camberwell, Vic, 138 pp. Ling, S.D., 2008: Range expansion of a habitat-modifying species leads to loss of taxonomic diversity: a new and impoverished reef state. Oecologia, 156(4), 883-894. Ling, S.D., C.R. Johnson, S. Frusher, C.K. King, 2008: Reproductive potential of a marine ecosystem engineer at the edge of a newly expanded range. Global Change Biology, 14(4), 907-915. Ling, S.D., C.R. Johnson, K. Ridgway, A.J. Hobday, M. Haddon, 2009: Climate-driven range extension of a sea urchin: inferring future trends by analysis of recent population dynamics. Global Change Biology, 15(3), 719- 731. Ling, N., 2010: Socio-economic drivers of freshwater fish declines in a changing climate: a New Zealand perspective. Journal of Fish Biology, 77(8), 1983-1992. Linnenluecke, M.K., A. Stathakis, A. Griffiths, 2011: Firm relocation as adaptive response to climate change and weather extremes. Global Environmental Change, 21(1), 123-133. Loechel, B., J.H. Hodgkinson, K. Moffat, S. Crimp, A. Littleboy, M. Howden, 2010: Goldfields-Esperance Regional Mining Climate Vulnerability Workshop: Report on workshop outcomes. CSIRO, Canberra, 42 pp. Loechel, B., J. Hodgkinson, K. Moffat, 2013: Climate Change Adaptation in Australian mining communities: comparing mining company and local government views and activities. Climatic Change, advance online, DOI: 10.1007/s10584-013-0721-8. Lough, J.M., 2008: Shifting climate zones for Australia's tropical marine ecosystems. Geophysical Research Letters, 35(14). Lough, J.M., A.J. Hobday, 2011: Observed climate change in Australian marine and freshwater environments. Marine and Freshwater Research, 62(9), 984-999. Loughnan, M.E., N. Nicholls, N.J. Tapper, 2010: The effects of summer temperature, age and socioeconomic circumstance on Acute Myocardial Infarction admissions in Melbourne, Australia. International Journal of Health Geographics, 210(9), 41. Lovelock, C.E., G. Skilleter, N. Saintilan, 2009: Tidal Wetlands and Climate Change. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 32 pp. Low Choy, D., S. Serrao-Neumann, F. Crick, G. Schuch, M. Sano, R.v. Staden, O. Sahin, B. Harman, S. Baum, 2012: Adaptation Options for Human Settlements in South East Queensland Main Report. South East Queensland Climate Adaptation Research Initiative, Griffith University, Nathan, Qld, 210 pp. Lucas, C., K.J. Hennessy, J.M. Bathols, 2007: Bushfire weather in Southeast Australia: recent trends and projected climate change impacts. Bushfire Cooperative Research Centre, Melbourne, 84 pp. Lundquist, C.J., D. Ramsay, R. Bell, A. Swales, S. Kerr, 2011: Predicted impacts of climate change on New Zealand's biodiversity. Pacific Conservation Biology, 17(3), 179-191. Lunt, I.D., M. Byrne, J.J. Hellmann, N.J. Mitchell, S.T. Garnett, M.W. Hayward, T.G. Martin, E. McDonald- Madden, S.E. Williams, K.K. Zander, 2013: Using assisted colonisation to conserve biodiversity and restore ecosystem function under climate change. Biological Conservation, 157, 172-177. Luo, Q., W. Bellotti, M. Williams, E. Wang, 2009: Adaptation to climate change of wheat growing in South Australia: Analysis of management and breeding strategies. Agriculture, Ecosystems & Environment, 129(1-3), 261-267. LVRC, 2012: Lockyer Valley Relocation Policy. Lockyer Valley Regional Council, Gatton, Qld, 13 pp. Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 LWF, 2010: Report of the Land and Water Forum: A Fresh Start for Freshwater. Land and Water Forum, Wellington, 85 pp. Mac Nally, R., G. Horrocks, H. Lada, P.S. Lake, J.R. Thomson, A.C. Taylor, 2009: Distribution of anuran amphibians in massively altered landscapes in south-eastern Australia: effects of climate change in an aridifying region. Global Ecology and Biogeography, 18(5), 575-585. Macintosh, A., 2013: Coastal climate hazards and urban planning: how planning responses can lead to maladaptation. Mitigation and Adaptation Strategies for Global Change, 18(7), 1035-1055. Macintosh, A., A. Foerster, J. McDonald, 2013: Limp, leap or learn? Developing legal frameworks for climate change adaptation planning in Australia. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 277 pp. Mackey, B.G., J.E.M. Watson, G. Hope, S. Gilmore, 2008: Climate Change, Biodiversity Conservation, and the Role of Protected Areas: An Australian Perspective. Biodiversity, 9, 11-18. Madin, E.M.P., N.C. Ban, Z.A. Doubleday, T.H. Holmes, G.T. Pecl, F. Smith, 2012: Socio-economic and management implications of range-shifting species in marine systems. Global Environmental Change-Human and Policy Dimensions, 22(1), 137-146. Maggini, R., H. Kujala, M.F.J. Taylor, J.R. Lee, H.P. Possingham, B.A. Wintle, R.A. Fuller, 2013: Protecting and restoring habitat to help Australia s threatened species adapt to climate change. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 53 pp. Maloney, S.K., C.F. Forbes, 2011: What effect will a few degrees of climate change have on human heat balance? Implications for human activity. International Journal of Biometeorology, 55(2), 147-160. Mankad, A., S. Tapsuwan, 2011: Review of socio-economic drivers of community acceptance and adoption of decentralised water systems. Journal of Environmental Management, 92(3), 380-391. Mapstone, B.D., P. Appleford, K. Broderick, R. Connolly, J. Higgins, A. Hobday, T.P. Hughes, P.A. Marshall, J. McDonald, M. Waschka, 2010: National Climate Change Adaptation Research Plan for Marine Biodiversity and Resources. National Climate Change Adaptation Research Facility, Gold Coast, Qld, 68 pp. Marcar, N.E., R.G. Benyon, P.J. Polglase, K.I. Paul, S. Theiveyanathan, L. Zhang, 2006: Predicting the hydrological impacts of bushfire and climate change in forested catchments of the River Murray uplands: a review. CSIRO Water for a Healthy Country National Research Flagship, Canberra, 38 pp. Marshall, P., J. Johnson, 2007: The Great Barrier Reef and climate change: vulnerability and management implications. Chapter 24. In: Climate change and the Great Barrier Reef. A vulnerability assessment. [Johnson, J., Marshall, P. (eds.)]. Great Barrier Reef Marine Park Authority and Australian Greenhouse Office, Townsville, Australia, pp. 745-771. Marshall, N.A., 2010: Understanding social resilience to climate variability in primary enterprises and industries. Global Environmental Change, 20(1), 36-43. Maru, Y.T., J. Langridge, B.B. Lin, 2011: Current and potential applications of typologies in vulnerability assessments and adaptation science. CSIRO Climate Adaptation Flagship Working paper No. 7. CSIRO, Melbourne, 25 pp. Maru, Y.T., V. Chewings, A. Sparrow, 2012: Climate change adaptation, energy futures and carbon economies in remote Australia: a review of the current literature, research and policy. CRC-REP Working Paper CW005. Ninti One Ltd, Alice Springs, 91 pp. Mason, M., K. Haynes, 2010: Adaptation Lessons from Cyclone Tracy. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 82 pp. Mason, M., K. Haynes, G. Walker, 2013: Cyclone Tracy and the road to improving wind resistant design. In: Natural Disasters and Adaptation to Climate Change [Boulter, S., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 137-149. Matusik, G., K.X. Ruthrof, N.C. Brouwers, B. Dell, G.S.J. Hardy, 2013: Sudden forest canopy collapse corresponding with extreme drought and heat in a mediterranean-type euclaypt forest in southwestern Australia. European Journal of Forest Research, 132, 497-510. Maunsell, CSIRO, 2008: Impacts of climate change on infrastructure in Australia and CGE model inputs. Report prepared by Maunsell Australia Pty Ltd in association with CSIRO Sustainable Ecosystems. Garnaut Climate Change Review, Canberra, 133 pp. Mayor, K., R.S.J. Tol, 2007: The impact of the UK aviation tax on carbon dioxide emissions and visitor numbers. Transport Policy, 14(6), 507-513. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McAdam, J., 2010: Disappearing states , statelessness and the boundaries of international law. In: Climate Change and Displacement [McAdam, J. (eds.)]. Hart Publishing, Oxford, UK, pp. 105-130. McAlpine, C.A., J. Syktus, R.C. Deo, P.J. Lawrence, H.A. McGowan, I.G. Watterson, S.R. Phinn, 2007: Modeling the impact of historical land cover change on Australia's regional climate. Geophysical Research Letters, 34(22), L22711. McAlpine, C.A., J. Syktus, J.G. Ryan, R.C. Deo, G.M. McKeon, H.A. McGowan, S.R. Phinn, 2009: A continent under stress: interactions, feedbacks and risks associated with impact of modified land cover on Australia's climate. Global Change Biology, 15(9), 2206-2223. McBride, J., N. Nicholls, 1983: Seasonal relationships between Australian rainfall and the Southern Oscillation. Monthly Weather Review, 111(10), 1998-2004. McBride, J., 2012: The Estimated Cost of Tropical Cyclone Impacts in Western Australia. Technical Report for the Indian Ocean Climate Initiative (IOCI) Stage 3. Project 2.2: Tropical Cyclones in the North West. Indian Ocean Climate Initiative (Government of Western Australia, CSIRO, Bureau of Meteorology), Perth, 53 pp. McCallum, J.L., R.S. Crosbie, G.R. Walker, W.R. Dawes, 2010: Impacts of climate change on groundwater in Australia: a sensitivity analysis of recharge. Hydrogeology Journal, 18(7), 1625-1638. McCleave, J., S. Espiner, K. Booth, 2006: The New Zealand people-park relationship: An exploratory model. Society and Natural Resources, 19(6), 547-561. McDonald, J., 2010: Paying the Price of Climate Change Adaptation: compensation for climate change impacts. In: Adaptation to Climate Change: Law and Policy [Bonyhady, T., Macintosh, A., McDonald, J. (eds.)]. Federation Press, Annandale, NSW, pp. 234-264. McDonald, J., 2011: The role of law in adapting to climate change. Wiley Interdisciplinary Reviews: Climate Change, 2, 283 295. McDonald, J., 2013: Creating legislative frameworks for adaptation. In: Climate Adaptation Futures [Palutikof, J., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 126-132. McGlone, M., S. Walker, R. Hay, J. Christie, 2010: Climate change, natural systems and their conservation in New Zealand. In: Climate change adaptation in New Zealand: Future scenarios and some sectoral perspectives [Nottage, R., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 82-99. McGlone, M.S., S. Walker, 2011: Potential effects of climate change on New Zealand s terrestrial biodiversity and policy recommendations for mitigation, adaptation and research. Science for Conservation 312. Department of Conservation, Wellington, 77 pp. McHenry, M.P., 2009: Agricultural bio-char production, renewable energy generation and farm carbon sequestration in Western Australia: Certainty, uncertainty and risk. Agriculture, Ecosystems & Environment, 129(1-3), 1-7. McInnes, K.L., D.J. Abbs, J.A. Bathols, 2005: Climate Change in Eastern Victoria. Stage 1 Report: The Effect of Climate Change on Coastal Wind and Weather Patterns. Gippsland Coastal Board, Melbourne, Vic, 26 pp. McInnes, K.L., I. Macadam, G.D. Hubbert, J.G. O'Grady, 2009: A modelling approach for estimating the frequency of sea level extremes and the impact of climate change in southeast Australia. Natural Hazards, 51(1), 115-137. McInnes, K.L., T.A. Erwin, J.M. Bathols, 2011a: Global Climate Model projected changes in 10 m wind speed and direction due to anthropogenic climate change. Atmospheric Science Letters, 12(4), 325-333. McInnes, K.L., I. Macadam, G.D. Hubbert, J.G. O'Grady, 2011b: An assessment of current and future vulnerability to coastal inundation due to sea level extremes in Victoria, southeast Australia. International Journal of Climatology, 33(1), 33-47. McInnes, K.L., J.G. O Grady, M. Hemer, I. Macadam, D.J. Abbs, 2012: Climate Futures for Tasmania: Extreme Sea Levels along Tasmania s Coast and the Impact of Climate Change. Antarctic Climate and Ecosystems CRC, 44 pp. McIntyre, S., 2011: Ecological and anthropomorphic factors permitting low-risk assisted colonization in temperate grassy woodlands. Biological Conservation, 144(6), 1781-1789. McIntyre-Tamwoy, S., A. Buhrich, 2011: Lost in the wash: predicting the impact of losing Aboriginal coastal sites in Australia. International Journal of Climate Change: Impacts and Responses, 3(1), 53-66. McKechnie, A.E., B.O. Wolf, 2010: Climate change increases the likelihood of catastrophic avian mortality events during extreme heat waves. Biology Letters, 6(2), 253-256. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McKeon, G.M., G.S. Stone, J.I. Syktus, J.O. Carter, N.R. Floof, D.G. Ahrens, D.N. Bruget, C.R. Chilcott, D.H. Cobon, R.A. Cowley, S.J. Crimp, G.W. Fraser, S.M. Howden, P.W. Johnston, J.G. Ryan, C.J. Stokes, K.A. Day, 2009: Climate change impacts on Australia's rangeland livestock carrying capacity: A review of challenges. Land & Water Australia, Canberra, 69 pp. McKerchar, A., B. Mullan, 2004: Seasonal inflow distributions for New Zealand hydroelectric power stations. NIWA Client Report: CHC2004-131. Ministry of Economic Development, Wellington, NZ, 12 pp. McKerchar, A.I., J.A. Renwick, J. Schmidt, 2010: Diminishing streamflows on the east coast of the South Island New Zealand and linkage to climate variability and change. Journal of Hydrology (NZ), 49, 1-14. Mcleod, D.J., G.M. Hallegraeff, G.W. Hosie, A.J. Richardson, 2012: Climate-driven range expansion of the red-tide dinoflagellate Noctiluca scintillans into the Southern Ocean. Journal of Plankton Research, 34(4), 332-337. McManus, P., J. Walmsley, N. Argent, S. Baumc, L. Bourked, J. Martine, B. Pritchard, T. Sorensen, 2012: Rural Community and Rural Resilience: What is important to farmers in keeping their country towns alive? Journal of Rural Studies, 28(1), 20-29. McMichael, A., H.J. Weaver, H. Berry, P. Beggs, B.J. Currie, J. Higgins, B. Kelly, J. McDonald, S. Tong, 2009: National climate change adaptation plan for human health. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 64 pp. McMillan, H., B. Jackson, S. Poyck, 2010: Flood Risk Under Climate Change. NIWA Report CHC2010-033. Ministry for Agriculture and Forestry, Wellington, 53 pp. McMillan, H., M. Duncan, G. Smart, J. Sturman, S. Poyck, S. Reese, A. Tait, E. Hreinsson, J. Walsh, 2012: The urban impacts toolbox: An example of modelling the effect of climate change and sea level rise on future flooding. Weather and Climate, 32(2), 21-39. McNamara, K.E., S.G. Smithers, W. R., K. Parnell, 2011: Limits to Climate Change Adaptation for Low-Lying Communities in the Torres Strait. Final report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 93 pp. McVicar, T.R., T.G. Van Niel, L.T. Li, M.L. Roderick, D.P. Rayner, L. Ricciardulli, R.J. Donohue, 2008: Wind speed climatology and trends for Australia, 1975-2006: Capturing the stilling phenomenon and comparison with near-surface reanalysis output. Geophysical Research Letters, 35(20), L035627. McVicar, T.R., R.J. Donohue, A.P. O'Grady, L.T. Li, 2010: The Effects of Climatic Changes on Plant Physiological and Catchment Ecohydrological Processes in the High-Rainfall Catchments of the Murray-Darling Basin: A Scoping Study. CSIRO Water for a Healthy Country National Research Flagship. Murray-Darling Basin Authority, Canberra, 95 pp. MDBA, 2011: Proposed Basin Plan. Publication 192/11. Draft plan prepared for the Commonwealth of Australia. Murray-Darling Basin Authority, Canberra, 226 pp. MDBA, 2012: Proposed Basin Plan - A Revised Draft. Publication 57/12. Murray-Darling Basin Authority, Canberra, 235 pp. MDBC, 2007: Impact of the 2003 Alpine Bushfires on Streamflow: Broadscale Water Yield Assessment. Murray Darling Basin Commission, Canberra, 122 pp. Measham, T., B. Preston, T. Smith, C. Brooke, R. Gorddard, G. Withycombe, C. Morrison, 2011: Adapting to climate change through local municipal planning: barriers and challenges. Mitigation and Adaptation Strategies for Global Change, 16(8), 889-909. MED, 2011: New Zealand's Energy Outlook 2011. Ministry of Economic Development, Wellington, 12 pp. Medlyn, B.E., R.A. Duursma, M.J.B. Zeppel, 2011a: Forest productivity under climate change: a checklist for evaluating model studies. Wiley Interdisciplinary Reviews: Climate Change, 2(3), 332-355. Medlyn, B.E., M. Zeppel, N.C. Brouwers, K. Howard, E. O Gara, G. Hardy, T. Lyons, L. Li, B. Evans, 2011b: Biophysical impacts of climate change on Australia's forests. Contribution of Work Package 2 to the Forest Vulnerability Assessment. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 189 pp. Meehl, G.A., T.F. Stocker, W.D. Collins, P. Friedlingstein, A.T. Gaye, J.M. Gregory, A. Kitoh, R. Knutti, J.M. Murphy, A. Noda, S.C.B. Raper, I.G. Watterson, A.J. Weaver, Z.-C. Zhao, 2007: Global Climate Projections. In: Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., Miller, H.L. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 747-845. Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Melbourne Water, 2010: Melbourne Water Responding to a Changing Climate. Melbourne Water, Melbourne, 19 pp. Melloy, P., G. Hollaway, J. Luck, R. Norton, E. Aitken, S. Chakraborty, 2010: Production and fitness of Fusarium pseudograminearum inoculum at elevated carbon dioxide in FACE. Global Change Biology, 16(12), 3363- 3373. Memon, A., P. Skelton, 2007: Institutional Arrangements and Planning Practices to Allocate Freshwater Resources in New Zealand: A Way Forward. New Zealand Journal of Environmental Law, 11, 241-277. Memon, A., B. Painter, E. Weber, 2010: Enhancing Potential for Integrated Catchment Management in New Zealand: A Multi-scalar, Strategic Perspective. Australasian Journal of Environmental Management, 17(1), 35- 44. Mendham, E., A. Curtis, 2010: Taking Over the Reins: Trends and Impacts of Changes in Rural Property Ownership. Society & Natural Resources, 23(7), 653-668. Menendez, M., P.L. Woodworth, 2010: Changes in extreme high water levels based on a quasi-global tide-gauge data set. Journal of Geophysical Research-Oceans, 115(C10), C10011. Meyssignac, B., A. Cazenave, 2012: Sea level: A review of present-day and recent-past changes and variability. 58(0), 96-109. MfE, 2007a: Environment New Zealand 2007. Report ME 847. Ministry for the Environment, Wellington, 456 pp. MfE, 2007b: Consultation with M ori on climate change: Hui Report. Report ME 830. Ministry for the Environment, Wellington, New Zealand, 135 pp. MfE, 2008a: Preparing for climate change: A guide for local government in New Zealand. Ministry for the Environment, Wellington, 44 pp. MfE, 2008b: Climate Change Effects and Impacts Assessment: A Guidance Manual for Local Government in New Zealand (2nd edition). Report ME870, prepared by B. Mullan, D. Wratt, S. Dean, M. Hollis, S. Allan, T. Williams, G. Kenny. Ministry for the Environment, Wellington, 167 pp. MfE, 2008c: Climate Change and Long-term Council Community Planning. Ministry for the Environment, Wellington, 22 pp. MfE, 2008d: Coastal Hazards and Climate Change: A Guidance Manual for Local Government in New Zealand (2nd edition). Revised by Ramsay, D., and Bell, R. (NIWA). Ministry for the Environment, Wellington, 139 pp. MfE, 2010a: Legally Protected Conservation Land in New Zealand: Environmental Snapshot April 2010. INFO 492. Ministry for the Environment, Wellington, 7 pp. MfE, 2010b: Tools for Estimating the Effects of Climate Change on Flood Flow: A Guidance Manual for Local Government in New Zealand. Report ME1013. New Zealand Ministry of Environment, Wellington, New Zealand, 63 pp. MfE, 2011: Freshwater Management 2011. National Policy Statement. Ministry for the Environment, Wellington, 12 pp. MfE, 2013: New Zealand s Greenhouse Gas Inventory 1990 2011. Publication ME 1113. Ministry for the Environment, Wellington, 463 pp. Milfont, T., 2012: The interplay between information level, perceived efficacy and concern about global warming and climate change: A one-year longitudinal study. Risk Analysis, 32(6), 1003-1020. Milfont, T., N. Harré, C. Sibley, J. Duckitt, 2012: The climate change dilemma: Examining the association between parental status and political party support. Journal of Applied Social Pscyhology, 42(10), 2386-2410. Miller, K.J., A.A. Rowden, A. Williams, V. Haussermann, 2011: Out of Their Depth? Isolated Deep Populations of the Cosmopolitan Coral Desmophyllum dianthus May Be Highly Vulnerable to Environmental Change. Plos One, 6(5). Miller, G.M., S.A. Watson, J.M. Donelson, M.I. McCormick, P.L. Munday, 2012: Parental environment mediates impacts of increased carbon dioxide on a coral reef fish. Nature Climate Change, 2(12), 858-861. Mills, R., 2009: Way Ahead. Rio Tinto Review, March 2009. Rio Tinto, Melbourne, Vic, 2 pp. Min, S.-K., W. Cai, P. Whetton, 2013: Influence of climate variability on seasonal extremes over Australia. Journal of Geophysical Research: Atmospheres, 118(2), 643-654. Minister of Conservation, 2010: New Zealand Coastal Policy Statement. Ministry of Conservation, Wellington, 31 pp. Mitchell, N.J., F.W. Allendorf, S.N. Keall, C.H. Daugherty, N.J. Nelson, 2010: Demographic effects of temperature-dependent sex determination: will tuatara survive global warming? Global Change Biology, 16(1), 60-72. Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Moise, A.F., D.A. Hudson, 2008: Probabilistic predictions of climate change for Australia and southern Africa using the reliability ensemble average of IPCC CMIP3 model simulations. Journal of Geophysical Research- Atmospheres, 113(D15), D009250. Moon, J.W., S.-H. Han, 2011: Thermostat strategies impact on energy consumption in residential buildings. Energy and Buildings, 43, 338-346. Moore, J.A.Y., L.M. Bellchambers, M.R. Depczynski, R.D. Evans, S.N. Evans, S.N. Field, K.J. Friedman, J.P. Gilmour, T.H. Holmes, R. Middlebrook, B.T. Radford, T. Ridgway, G. Shedrawi, H. Taylor, D. Thomson, S.K. Wilson, 2012: Unprecedented mass bleaching and loss of coral across 12° of latitude in Western Australia in 2010-2011. Plos One, 7(12), e51807. Moore, A.D., A. Ghahramani, 2013: Climate change and broadacre livestock production across southern Australia. 1. Impacts of climate change on pasture and livestock productivity, and on sustainable levels of profitability. Global Change Biology, 19(5), 1440-1455. Morrison, C., C.M. Pickering, 2011: Climate change adaptation in the Australian Alps: impacts, strategies, limits and management. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 84 pp. Morrongiello, J.R., S.J. Beatty, J.C. Bennett, D.A. Crook, D.N.E.N. Ikedife, M.J. Kennard, A. Kerezsy, M. Lintermans, D.G. McNeil, B.J. Pusey, T. Rayner, 2011: Climate change and its implications for Australia's freshwater fish. Marine and Freshwater Research, 62, 1082-1098. Mortimer, E., A. Bergin, R. Carter, 2011: Sharing risk: financing Australia s disaster resilience. Australian Strategic Policy Institute, Barton, ACT, 24 pp. Mukheibir, P., N. Kuruppu, A. Gero, J. Herriman, 2013: Cross-Scale Barriers to Climate Change Adaptation in Local Government, Australia. Final report. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 101 pp. Mulet-Marquis, S., J.R. Fairweather, 2008: Rural Population and Farm Labour Change. Lincoln University, Lincoln, New Zealand, 35 pp. Mullan, B., 1995: On the linearity and stability of Southern Oscillation-climate relationships for New Zealand. International Journal of Climatology, 15(12), 1365-1386. Mullan, B., M.J. Salinger, C.S. Thompson, A.S. Porteous, 2001: The New Zealand climate: present and future. In: The Effects of climate change and variation in New Zealand: An assessment using the Climpacts system [Warrick, R.A., G.J. Kenny and J.J. Harman (eds.)]. IGCI, Waikato, Hamilton, pp. 11-31. Mullan, B., A. Porteous, D. Wratt, M. Hollis, 2005: Changes in drought risk with climate change. NIWA Reort WLG2005-23. Ministry for the Environment, Wellington, 58 pp. Mullan, B., S.J. Stuart, M.G. Hadfield, M.J. Smith, 2010: Report on the Review of NIWA s Seven-Station Temperature Series. NIWA Information Series No. 78. National Institute of Water and Atmospheric Research (NIWA), Wellington, 175 pp. Mullan, B., T. Carey-Smith, G.M. Griffiths, A. Sood, 2011: Scenarios of storminess and regional wind extremes under climate change. NIWA client report WLG2010-31. Ministry of Agriculture and Forestry, Wellington, 80 pp. Munday, P.L., N.E. Crawley, G.E. Nilsson, 2009: Interacting effects of elevated temperature and ocean acidification on the aerobic performance of coral reef fishes. Marine Ecology-Progress Series, 388, 235-242. Murphy, H., A. Liedloff, R.J. Williams, K.J. Williams, M. Dunlop, 2012: Queensland's biodiversity under climate change: terrestrial ecosystems. CSIRO Climate Adaptation Flagship Working Paper No. 12C. CSIRO, Canberra, 120 pp. Mustelin, J., N. Kuruppu, A.M. Kramer, J. Daron, K. de Bruin, A.G. Noriega, 2013: Climate Adaptation Research for the Next Generation. Climate and Development, advance online, DOI: 10.1080/17565529.2013.812953. Naish, S., K. Mengersen, W. Hu, S. Tong, 2013: Forecasting the futrue risk of Barmah Forest Virus disease under climate change sceanrios in Queensland, Australia. Plos One, 8(5), e62843. Nana, G., F. Stokes, W. Molano, 2011a: The Asset base, Income, Expenditure and GDP of the 2010 M ori Economy. Report for the M ori economic taskforce. Business and Economic Research Ltd, Wellington, New Zealand, 44 pp. Nana, G., F. Stokes, W. Molano, 2011b: The M ori economy, science and innovation scenarios. Report for the M ori economic taskforce. Business and Economic Research Ltd, Wellington, New Zealand, 42 pp. NDIR, 2011: Inquiry into flood insurance and related matters. Natural Disaster Insurance Review, Canberra, 149 pp. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Nearing, M.A., F.F. Pruski, M.R. O'Neal, 2004: Expected climate change impacts on soil erosion rates: a review. Journal of Soil and Water Conservation, 59(1), 43-50. Nelson, R., M. Howden, M.S. Smith, 2008: Using adaptive governance to rethink the way science supports Australian drought policy. Environmental Science & Policy, 11(7), 588-601. Nelson, R., P. Kokic, S. Crimp, P. Martin, H. Meinke, S.M. Howden, P. de Voil, U. Nidumolu, 2010: The vulnerability of Australian rural communities to climate variability and change: Part II-Integrating impacts with adaptive capacity. Environmental Science and Policy, 13(1), 18-27. Neuheimer, A.B., R.E. Thresher, J.M. Lyle, J.M. Semmens, 2011: Tolerance limit for fish growth exceeded by warming waters. Nature Climate Change, 1(2), 110-113. Newton, P.C.D., V. Allard, R.A. Carran, M. Lieffering, 2006: Impacts of elevated CO2 on a grassland grazed by sheep: the New Zealand FACE experiment. In: Managed Ecosystems and CO2: Case Studies, Processes, and Perspectives [Nosberger, J., Long, S.P., Norby, R.J., Stitt, M., Hendrey, G.R., Blum, H. (eds.)]. Springer- Verlag, Berlin, Germany, pp. 157-171. Newton, P.C.D., M. Lieffering, A.J. Parsons, S.C. Brock, P.W. Theobald, C.L. Hunt, D. Luo, M.J. Hovenden, 2013: Selective grazing modifies previously anticipated responses of plant community composition to elevated CO2 in a temperate grassland. Global Change Biology, advance online, DOI: 10.1111/gcb.12301. NFRAG, 2008: National Flood Risk Advisory Group: Flood risk management in Australia. Australian Journal of Emergency Management, 23, 21-27. Nguyen, M., X. Wang, D. Chen, 2010: An investigation of extreme heatwave events and their effects on building and infrastructure. CSIRO, Canberra, 113 pp. Nicholls, N., 2006: Detecting and attributing Australian climate change: a review. Australian Meteorological Magazine, 55(3), 199-211. Nicholls, N., D. Collins, 2006: Observed climate change in Australia over the past century. Energy & Environment, 17(1), 1-12. Nicholls, N., 2010: Local and remote causes of the southern Australian autumn-winter rainfall decline, 1958-2007. Climate Dynamics, 34(6), 835-845. Nicholls, N., P. Uotila, L. Alexander, 2010: Synoptic influences on seasonal, interannual and decadal temperature variations in Melbourne, Australia. International Journal of Climatology, 30(9), 1372-1381. Nicholls, N., 2011: Comments on Influence of Location, Population, and Climate on Building Damage and Fatalities due to Australian Bushfire: 1925 2009 . Weather, Climate, and Society, 3(1), 61-62. Nidumolu, U., S. Crimp, D. Gobbett, A. Laing, M. Howden, S. Little, 2011: Heat stress in dairy cattle in northern Victoria: responses to a changing climate. Climate Adaptation Flagship Working paper No. 10. CSIRO, Canberra, 72 pp. Nilsson, G.E., D.L. Dixson, P. Domenici, M.I. McCormick, C. Sorensen, S.A. Watson, P.L. Munday, 2012: Near- future carbon dioxide levels alter fish behaviour by interfering with neurotransmitter function. Nature Climate Change, 2(3), 201-204. NIWA, MWH, GNS Science, BRANZ, 2012: Impacts of Climate Change on Urban Infrastructure and the Built Environment: Toolbox Handbook. National Institute of Water and Atmospheric Research (NIWA), Wellington, 34 pp. NNTT, 2013: National Native Title Tribunal Dataset: Determinations of Native Title, as per s. 192 Native Title Act 1993 (Cwlth), National Native Title Register - boundaries and core attributes (online dataset, accessed 31 August 2013, at www.ga.gov.au/meta/ANZCW0703011416.html). Geoscience Australia, Canberra. Norman, B., 2009: Principles for an intergovernmental agreement for coastal planning and climate change in Australia. Habitat International, 33(3), 293-299. Norman, B., 2010: A Low Carbon and Resilient Urban Future. Department of Climate Change and Energy Efficiency, Canberra, 60 pp. Norman, B., W. Steffen, R. Webb, A. Capon, W. Maher, C. Woodroffe, K. Rogers, R. Tanton, Y. Vidyattama, J. Lavis, H. Sinclair, B. Weir, 2012: South East Coastal Adaptation (SECA): Coastal urban climate futures in SE Australia from Wollongong to Lakes Entrance. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 130 pp. Norman-Lopez, A., S. Pascoe, A.J. Hobday, 2011: Potential economic impacts of climate change on Australian fisheries and the need for adaptive management. Climate Change Economics, 2, 209-235. NRC, 2012: Framework for assessing and recommending upgraded catchment action plans. Natural Resources Commission, New South Wales Government, Sydney, 32 pp. Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 NRMMC, 2010: National Feral Camel Action Plan: A national strategy for the management of feral camels in Australia. Vertebrate Pests Committee, Natural Resource Management Ministerial Council. Department of Sustainability, Environment, Water, Population and Communities, Canberra, 72 pp. Nursey-Bray, M., D. Fergie, V. Arbon, L. Rigney, R. Palmer, J. Tibby, N. Harvey, L. Hackworth, 2013: Community Based Adaptation to Climate Change: The Arabana. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 133 pp. NWC, 2009: Australian Water Reform 2009: Second Biennial Assessment of Progress in Implementation of the National Water Initiative. National Water Commission, Canberra, 288 pp. NWC, 2010: The Impacts of Water Trading in the Southern Murray-Darling Basin: An Economic, Social and Environmental Assessment. National Water Commission, Canberra, 237 pp. NWC, 2011: The National Water Initiative - Securing Australia's Water Future: 2011 Assessment. National Water Commission, Canberra, 354 pp. NZIER, 2003: M ori Economic Development: Te Ohanga Whakaketanga M ori. New Zealand Institute of Economic Research, Wellington, New Zealand, 116 pp. O'Connell, M., R. Hargreaves, 2004: Climate change adaptation: guidance on adapting New Zealand s built environment for the impacts of climate change. Study Report No.130. Building Research Association of New Zealand (BRANZ), Wellington, 44 pp. O'Donnell, L., 2007: Climate Change. An analysis of the policy considerations for climate change for the Review of the Canterbury Regional Policy Statement. R07/4. Environment Canterbury, Christchurch, 75 pp. O'Leary, G., B. Christy, A. Weeks, C. Beverly, J. Nuttall, P. Riffkin, G. Fitzgerald, R. Norton, 2010: Modelling the growth and grain yield response of wheat crops under Free Air Carbon Dioxide Enrichment at Horsham and throughout Victoria, Australia. In: Food Security from Sustainable Agriculture. 15th Agronomy Conference, 15- 18 November 2010, Lincoln, New Zealand [Dove, H., Culvenor, R.A. (eds.)], Australian Society of Agronomy, Gosford, NSW. O'Neill, S.J., J. Handmer, 2012: Responding to bushfire risk: the need for transformative adaptation. Environmental Research Letters, 7(1), 014018. Odeh, I.O.A., D.K.Y. Tan, T. Ancev, 2011: Potential Suitability and Viability of Selected Biodiesel Crops in Australian Marginal Agricultural Lands Under Current and Future Climates. Bioenergy Research, 4(3), 165- 179. OECD, 2010: Sustainable Management of Water Resources in Agriculture. Organisation for Economic Co- Operation and Development, Paris, 122 pp. OECD, 2011: Dairy. In: Agricultural Outlook 2011 [OECD, FAO (eds.)]. OECD Publishing, Paris, pp. 159-173. OECD, 2013a: OECD Factbook 2013: Economic, Environmental and Social Statistics. OECD Publishing, Paris, 300 pp. OECD, 2013b: Water Security for Better Lives. OECD Studies on Water, OECD Publishing, Paris, 173 pp. Ogden, J., L. Basher, M.S. McGlone, 1998: Fire, forest regeneration and links with early human habitation: evidence from New Zealand. Annals of Botany, 81, 687-696. Oliver, R., I. Webster, 2011: Water for the environment. In: Water: Science and Solutions for Australia [Prosser, I.P. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 119-134. Oxford Economics, 2009: Valuing the effects of Great Barrier Reef Bleaching. Great Barrier Reef Foundation, Newstead, Qld, 102 pp. Packman, D., D. Ponter, T. Tutua-Nathan, 2001: Climate Change Working Paper: M ori Issues. New Zealand Climate Change Programme, Wellington, New Zealand, 18 pp. Palutikof, J., M. Parry, M. Stafford-Smith, A.J. Ash, S.L. Boulter, M. Waschka, 2013: The past, present and future of adaptation: setting the context and naming the challenges. In: Climate Adaptation Futures [Palutikof, J., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 3-29. Park, S.E., S. Crimp, N.G. Inman-Bamber, Y.L. Everingham, 2010: Sugarcane. In: Adapting Agriculture to Climate Change. Preparing Australian Agriculture, Forestry and Fisheries for the Future [Stokes, C., Howden, S. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 85-99. Park, S.E., S. Crimp, A.M. Dowd, P. Ashworth, E. Mendham, 2011: Understanding climate change adaptation as a national challenge. Working Paper No. 8. CSIRO Climate Adaptation Flagship, Canberra, 32 pp. Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Park, S.E., N.A. Marshall, E. Jakku, A.M. Dowd, S.M. Howden, E. Mendham, A. Fleming, 2012: Informing adaptation responses to climate change through theories of transformation. Global Environmental Change, 22(1), 115-126. Parker, D., C. Folland, A. Scaife, J. Knight, A. Colman, P. Baines, B. Dong, 2007: Decadal to multidecadal variability and the climate change background. Journal of Geophysical Research-Atmospheres, 112(D18), D18115. Parsons Brinkerhoff, 2009: Energy network infrastructure and the climate change challenge. Energy Networks Association, Canberra, 129 pp. Peace, R., 2001: Social exclusion: a concept in need of definition? Social Policy Journal of New Zealand, 16, 17-36. Pearce, A., M. Feng, 2007: Observations of warming on the Western Australian continental shelf. Marine and Freshwater Research, 58(10), 914-920. Pearce, H.G., G. Cameron, S.A.J. Anderson, M. Dudfield, 2008: An overview of fire management in New Zealand forestry. New Zealand Journal of Forestry, 53(3), 7-11. Pearce, H.G., J. Kerr, A. Clark, B. Mullan, D. Ackerley, T. Carey-Smith, E. Yang, 2011: Improved estimates of the effect of climate change on NZ fire danger. Scion Client Report No. 18087 in conjunction with NIWA, Wellington. New Zealand Fire Service, Wellington, 84 pp. Pearce, M., E. Willis, T. Jenkin, 2007: Aboriginal people's attitudes towards paying for water in a water-scarce region of Australia. Environment Development and Sustainability, 9(1), 21-32. Pearman, G., 2009: Climate change and security: Why is this on the agenda? In: Climate change and security. Planning for the future [Boston, J., Nel, P., Righarts, M. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, NZ, pp. 9-36. Peel, M.C., T.A. McMahon, B.L. Finlayson, 2004: Continental differences in the variability of annual runoff-update and reassessment. Journal of Hydrology, 295(1-4), 185-197. Pecl, G., S. Frusher, C. Gardner, M. Haward, A. Hobday, S. Jennings, M. Nursey-Bray, A. Punt, H. Revill, I. van Putten, 2009: The east coast Tasmanian rock lobster fishery vulnerability to climate change impacts and adaptation response options. Department of Climate Change, Canberra, 100 pp. Penny, G., D. King, 2010: Transfer of Renewable Energy Technology Systems to Rural M ori Communities. National Institute of Water and Atmospheric Research (NIWA), Auckland, 59 pp. Petheram, L., K.K. Zander, B.M. Campbell, C. High, N. Stacey, 2010: 'Strange changes': Indigenous perspectives of climate change and adaptation in NE Arnhem Land (Australia). Global Environmental Change, 20(4), 681- 692. Petheram, C., P. Rustomji, F. Chiew, W. Cai, T. McVicar, J. Vleeshouwer, T. Van Neil, L. Li, R. Donohue, R. Cresswell, J. Teng, J.-M. Perraud, 2012: Estimating the impact of projected climate change on runoff across the tropical savannas and semi-arid rangelands of northern Australia. Journal of Hydrometeorology, 13(2), 483 503. Petrone, K., J. Hughes, T. Van Niel, R. Silberstein, 2010: Streamflow decline in southwestern Australia, 1950- 2008. Geophysical Research Letters, 37, L043102. Pettit, C.J., C.M. Raymond, B.A. Bryan, H. Lewis, 2011: Identifying strengths and weaknesses of landscape visualisation for effective communication of future alternatives. Landscape and Urban Planning, 100(3), 231- 241. Pham, T.D., D.G. Simmons, R. Spurr, 2010: Climate change-induced economic impacts on tourism destinations: the case of Australia. Journal of Sustainable Tourism, 18(3), 449-473. Phelan, L., 2011: Managing climate risk: extreme weather events and the future of insurance in a climate-changed world. Australian Journal of Environmental Management, 18, 223-232. Pickering, C.M., W. Hill, K. Green, 2008: Vascular plant diversity and climate change in the alpine zone of the Snowy Mountains, Australia. Biodiversity and Conservation, 17(7), 1627-1644. Pickering, C.M., R.C. Buckley, 2010: Climate Response by the Ski Industry: The Shortcomings of Snowmaking for Australian Resorts. AMBIO: A Journal of the Human Environment, 39(5-6), 430-438. Pickering, C.M., J.G. Castley, M. Burtt, 2010: Skiing Less Often in a Warmer World: Attitudes of Tourists to Climate Change in an Australian Ski Resort. Geographical Research, 48(2), 137-147. Pitt, N.R., E.S. Poloczanska, A.J. Hobday, 2010: Climate-driven range changes in Tasmanian intertidal fauna. Marine and Freshwater Research, 61(9), 963-970. Pittock, J., L.J. Hansen, R. Abell, 2008: Running dry: freshwater biodiversity, protected areas and climate change. Biodiversity, 9(3/4), 30-38. Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pittock, J., C.M. Finlayson, 2011: Australia's Murray-Darling Basin: freshwater ecosystem conservation options in an era of climate change. Marine and Freshwater Research, 62(3), 232-243. Pittock, J., 2013: Lessons from adaptation to sustain freshwater environments in the Murray Darling Basin, Australia. Wiley Interdisciplinary Reviews: Climate Change, 4(5), 429-438. PMSEIC, 2010a: Australia and Food Security in a Changing World. Prime Minister s Science, Engineering and Innovation Council, Canberra, 90 pp. PMSEIC, 2010b: Challenges at Energy-Water-Carbon Intersections. Prime Minister s Science, Engineering and Innovation Council, Canberra, 88 pp. Polglase, P., A. Reeson, C. Hawkins, K. Paul, A. Siggins, J. Turner, D. Crawford, T. Jovanovic, T. Hobbs, K. Opie, J. Carwardine, A. Almeida, 2011: Opportunities for carbon forestry in Australia: Economic assessment and constraints to implementation. CSIRO, Canberra, 30 pp. Poloczanska, E.S., R.C. Babcock, A. Butler, A. Hobday, O. Hoegh-Guldberg, T.J. Kunz, R. Matear, D.A. Milton, T.A. Okey, A.J. Richardson, 2007: Climate change and Australian marine life. Oceanography and Marine Biology, 45, 407-478. Pomeroy, A., 1996: Relocating policies for sustainable agriculture under the umbrella of rural development in Australia, USA and New Zealand. New Zealand Geographer, 52(2), 84-87. Poot, J., 2009: Trans-Tasman Migration, Transnationalism and Economic Development in Australasia. Motu Working Paper 09-05. Motu Economic and Public Policy Research, Wellington, 21 pp. Post, D.A., J. Teng, F. Chiew, B. Wang, J. Vaze, S. Marvanek, S.W. Franks, E. Boegh, E. Blyth, D.M. Hannah, K.K. Yilmaz, 2011: Non-linearity of the runoff response across southeastern Australia to increases in global average temperature. In: Hydro-Climatology: Variability and Change. IAHS Red Book 344 [Franks, S.W., Boegh, E., Blyth, E., Hannah, D.M., Yilmaz, K.K. (eds.)]. International Association of Hydrological Sciences, Oxfordshire, UK, pp. 188-194. Post, D.A., F. Chiew, J. Teng, N.R. Viney, F.L.N. Ling, G. Harrington, R.S. Crosbie, B. Graham, S. Marvanek, R. McLoughlin, 2012: A robust methodology for conducting large-scale assessments of current and future water availability and use: a case study in Tasmania, Australia. Journal of Hydrology, 412-413, 233-245. Potter, N.J., F. Chiew, A.J. Frost, 2010: An assessment of the severity of recent reductions in rainfall and runoff in the Murray-Darling Basin. Journal of Hydrology, 381(1-2), 52-64. Potter, N.J., F. Chiew, 2011: An investigation into changes in climate characteristics causing the recent very low runoff in the southern Murray-Darling Basin using rainfall-runoff models. Water Resources Research, 47, W00G10. Power, S., F. Tseitkin, S. Torok, B. Lavery, R. Dahni, B. McAvaney, 1998: Australian temperature, Australian rainfall and the Southern Oscillation, 1910-1992: coherent variability and recent changes. Australian Meteorological Magazine, 47(2), 85-101. Power, S., B. Sadler, N. Nicholls, 2005: The Influence of Climate Science on Water Management in Western Australia: Lessons for Climate Scientists. Bulletin of the American Meteorological Society, 86(6), 839-844. Poyck, S., J. Hendrikx, H. McMillan, E. Hreinsson, R. Woods, 2011: Combined snow and streamflow modelling to estimate impacts of climate change on water resources in the Clutha River, New Zealand. Journal of Hydrology, 50(2), 293-312. Pratchett, M.S., L.K. Bay, P.C. Gehrke, J.D. Koehn, K. Osborne, R.L. Pressey, H.P.A. Sweatman, D. Wachenfeld, 2011: Contribution of climate change to degradation and loss of critical fish habitats in Australian marine and freshwater environments. Marine and Freshwater Research, 62(9), 1062-1081. Preston, B., R.N. Jones, 2008: Screening Climatic and Non-Climatic Risks to Australian Catchments. Geographical Research, 46(3), 258-274. Preston, B., T.F. Smith, C. Brooke, R. Gorddard, T. Measham, G. Withycombe, B. Beveridge, C. Morrison, K. McInnes, D. Abbs, 2008: Mapping climate change vulnerability in the Sydney Coastal Councils Group. Sydney Coastal Councils Group and Australian Government, Sydney, 124 pp. Preston, B., M. Stafford-Smith, 2009: Framing vulnerability and adaptive capacity assessment: Discussion paper. CSIRO Climate Adaptation Flagship Working paper No. 2. CSIRO, Canberra. Preston, B., C. Brooke, T. Measham, T. Smith, R. Gorddard, 2009: Igniting change in local government: lessons learned from a bushfire vulnerability assessment. Mitigation and Adaptation Strategies for Global Change, 14(3), 251-283. Preston, B., R. Westaway, E. Yuen, 2011: Climate adaptation planning in practice: an evaluation of adaptation plans from three developed nations. Mitigation and Adaptation Strategies for Global Change, 16(4), 407-438. Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pride, S., B. Frame, D. Gill, 2010: Futures Literacy in New Zealand. Journal of Futures Studies, 15(1), 135-144. Prober, S.M., M. Dunlop, 2011: Climate change: a cause for new biodiversity conservation objectives but let s not throw the baby out with the bathwater. Ecological Management and Restoration, 12, 2-3. Productivity Commission, 2009: Government Drought Support. Final Inquiry Report No. 46. Productivity Commission, Melbourne, 431 pp. Productivity Commission, 2012: Barriers to effective climate change adaptation. Inquiry Report No. 59. Productivity Commission, Canberra, 385 pp. Prosser, I.P., I.D. Rutherfurd, J.M. Olley, W.J. Young, P.J. Wallbrink, C.J. Moran, 2001: Large-scale patterns of erosion and sediment transport in rivers networks, with examples from Australia. Freshwater and Marine Research, 52(1), 81-99. Prowse, T.A.A., B.W. Brook, 2011: Climate change, variability and conservation impacts in Australia. Pacific Conservation Biology, 17(3), 168-178. Purdie, H., A. Mackintosh, W. Lawson, B. Anderson, U. Morgenstern, T. Chinn, P. Mayewski, 2011: Interannual variability in net accumulation on Tasman Glacier and its relationship with climate. Global and Planetary Change, 77(3-4), 142-152. PWC, 2011: Economic impact of Queensland s natural disasters. March 2011. PriceWaterhouseCoopers, Sydney, NSW, 6 pp. QFCI, 2012: Queensland Floods Commission of Inquiry Final Report. Queensland Floods Commission of Inquiry, Brisbane, Qld, 623 pp. QRC, 2011: Queensland Resources Council, submission to Queensland floods commision of inquiry. Queensland Resources Council, Brisbane, Qld, 43 pp. Queensland Government, 2011: Understanding Floods: Questions and Answers. Queensland Government, Brisbane, 36 pp. Quiggin, J., D. Adamson, P. Schrobback, S. Chambers, 2008: The Implications for Irrigation in the Murray-Darling Basin. Paper prepared for the Garnaut Climate Change Review. Garnaut Climate Change Review, Canberra, 59 pp. Quiggin, J., D. Adamson, S. Chambers, P. Schrobback, 2010: Climate Change, Uncertainty, and Adaptation: The Case of Irrigated Agriculture in the Murray Darling Basin in Australia. Canadian Journal of Agricultural Economics/Revue canadienne d'agroeconomie, 58(4), 531-554. Qureshi, E.M., M.A. Hanjra, J. Ward, 2013a: Impact of water scarcity in Australia on global food security in an era of climate change. Food Policy, 38(0), 136-145. Qureshi, E.M., S.M. Whitten, M. Mohammed, S. Marvanek, A. Elmahdi, 2013b: A biophysical and economic model of agriculture and water in the Murray-Darling Basin, Australia. Environmental Modelling & Software, 41(0), 98-106. QUT, 2010: Impacts and Adaptation Responses of Infrastructure and Communities to Extreme Events. The Southern Australian Experience of 2009. Prepared by Queensland University of Technology. National Climate Change Adaptation Research Facility, Gold Coast, Qld. Radcliffe, J.E., J.A. Baars, 1987: Temperate pasture production. In: Ecosystems of the World [Snaydon, R.W. (eds.)]. Elsevier, Amsterdam, pp. 7-17. Rafter, A., D.J. Abbs, 2009: An analysis of future changes in extreme rainfall over Australian regions based on GCM simulations and Extreme Value Analysis. CAWCR Research Letters (Centre for Australian Weather and Climate Research), Vol 3 [P. A. Sandery, T. Leeuwenburg, G. Wang, A. J. Hollis (eds.)], 44-49. Rajanayaka, C., J. Donaggio, H. McEwan, 2010: Update of Water Allocation Data and Estimate of Actual Water Use of Consented Takes - 2009-10. Report No H10002/3, October 2010. Ministry for the Environment, Wellington, 118 pp. Randall, A., T. Capon, T. Sanderson, D. Merrett, G. Hertzler, 2012: Making decisions under the risks and uncertainties of future climates. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 72 pp. Rau, G.H., E.L. McLeod, O. Hoegh-Guldberg, 2012: The need for new ocean conservation strategies in a high- carbon dioxide world. Nature Climate Change, 2(10), 720-724. Raymond, C.M., G. Brown, 2011: Assessing spatial associations between perceptions of landscape value and climate change risk for use in climate change planning. Climatic Change, 104(3-4), 653-678. Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Raymond, C.M., G. Brown, G.M. Robinson, 2011: The influence of place attachment, and moral and normative concerns on the conservation of native vegetation: A test of two behavioural models. Journal of Environmental Psychology, 31(4), 323-335. Raymond, C.M., J. Spoehr, 2013: The acceptability of climate change in agricultural communities: Comparing responses across variability and change. Journal of Environmental Management, 115, 69-77. RBA, 2006: Statement on Monetary Policy November 2006. Reserve Bank of Australia, Canberra, 59 pp. RBA, 2011: Statement on Monetary Policy. Box B: An update on the impact of natural disasters in Queensland. Reserve Bank of Australia, Canberra, 5 pp. Reijnders, L., M.A.J. Huijbregts, 2007: Life cycle greenhouse gas emissions, fossil fuel demand and solar energy conversion efficiency in European bioethanol production for automotive purposes. Journal of Cleaner Production, 15(18), 1806-1812. Reisinger, A., B. Mullan, M.R. Manning, D. Wratt, R. Nottage, 2010: Global and local climate change scenarios to support adaptation in New Zealand. In: Climate change adaptation in New Zealand: Future scenarios and some sectoral perspectives [Nottage, R., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 26-43. Reisinger, A., D. Wratt, S. Allan, H. Larsen, 2011: The role of local government in adapting to climate change: lessons from New Zealand. In: Climate change adaptation in developed nations. From theory to practice [Ford, J.D., Berrang-Ford, L. (eds.)]. Springer, Toronto, Canada, pp. 303-319. Reisinger, A., P. Havlik, K. Riahi, O. Vliet, M. Obersteiner, M. Herrero, 2012: Implications of alternative metrics for global mitigation costs and greenhouse gas emissions from agriculture. Climatic Change, 117(4), 677-690. Reisinger, A., J. Lawrence, G. Hart, R. Chapman, 2014: From coping to resilience: the role of managed retreat in highly developed coastal regions. In: Climate Change and the Coast: Building Resilient Communities [Glavovic, B., Kaye, R., Kelly, M., Travers, A. (eds.)]. CRC Press, London, www.crcpress.com/product/isbn/9780415464871. Ren, Z., X. Wang, Z. Chen, 2012: Energy efficient houses for heat stress mitigation: 2009 Melbourne heatwave scenarios. In: 10th International Healthy Building Conference, 8-12 July 2012 [International Society of Indoor Air Quality and Climate (eds.)], Brisbane, Qld. pp 4D.5. Renwick, J., P. Mladenov, J. Purdie, A. McKerchar, D. Jamieson, 2009: The effects of climate variability and change upon renewable electricity in New Zealand. In: Climate change adaptation in New Zealand: Future scenarios and some sectoral perspectives [Nottage, R., Wratt, D., Bornman, J., Jones, K. (eds.)]. New Zealand Climate Change Centre, Wellington, pp. 70-81. Reser, J.P., J.K. Swim, 2011: Adapting to and coping with the perceived threat and unfolding Impacts of global climate change. American Psychologist, 66(4), 277-289. Reser, J.P., S.A. Morrissey, M. Ellul, 2011: The threat of climate change: psychological response, adaptation, and impacts. In: Climate change and human well being. International and Cultural Psychology Series [Weissbecker, I. (eds.)]. Springer, New York, pp. 19-42. Reser, J.P., G.L. Bradley, M.C. Ellul, 2012a: Coping with climate change: Bringing psychological adaptation in from the cold. In: Handbook of the psychology of coping: Psychology of emotions, motivations and actions [Molinelli, B., Grimaldo, V. (eds.)]. Nova Science Publishers, New York, pp. 1-34. Reser, J.P., G.L. Bradley, A.I. Glendon, M.C. Ellul, R. Callaghan, 2012b: Public risk perceptions, understandings and responses to climate change and natural disasters in Australia, 2010 and 2011. Final report. Griffith University, School of Psychology, Gold Coast, Qld, 246 pp. Reser, J.P., G.L. Bradley, A.I. Glendon, M.C. Ellul, R. Callaghan, 2012c: Public Risk Perceptions, Understandings, and Responses to Climate Change and Natural Disasters in Australia and Great Britain. Griffith University, School of Psychology, Gold Coast, Qld, 298 pp. Reside, A.E., J. Van Der Wal, B.L. Phillips, L.P. Shoo, D.F. Rosauer, B.J. Anderson, J.A. Welbergen, C. Moritz, S. Ferrier, T.D. Harwood, K.J. Williams, Brendan Mackey, S. Hugh, S.E. Williams, 2013: Climate change refugia for terrestrial biodiversity: defining areas that promote species persistence and ecosystem resilience in the face of global climate change. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 230 pp. Rhodes, B.G., J.E. Fagan, K.S. Tan, 2012: Responding to a rapid climate shift - experiences from Melbourne, Australia. In: World Congress on Water, Climate and Energy, 13-18 May 2012 [International Water Association (eds.)], Dublin, Ireland. pp 8. Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Richardson, A.J., E.S. Poloczanska, 2009: Australia's Oceans. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 6 pp. Richardson, A.J., D. McKinnon, K.M. Swadling, 2009: Zooplankton. In: A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009 [Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (eds.)]. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, Publication 05/09, 15 pp. Rickards, L., S.M. Howden, 2012: Transformational adaptation: Agriculture and climate change. Crop and Pasture Science, 63(3), 240-250. Ridgway, K.R., 2007: Long-term trend and decadal variability of the southward penetration of the East Australian Current. Geophysical Research Letters, 34(13), L030393. Risbey, J.S., M.J. Pook, P.C. McIntosh, M.C. Wheeler, H.H. Hendon, 2009: On the Remote Drivers of Rainfall Variability in Australia. Monthly Weather Review, 137(10), 3233-3253. Ritchie, E.G., E.E. Bolitho, 2008: Australia's Savanna Herbivores: Bioclimatic Distributions and an Assessment of the Potential Impact of Regional Climate Change. Physiological and Biochemical Zoology, 81(6), 880-890. Rive, V., T. Weeks, 2011: Adaptation to climate change in New Zealand. In: Climate Change Law and Policy in New Zealand [Cameron, A. (eds.)]. LexisNexis, Wellington, pp. 345-392. Roberts, D.A., E.L. Johnston, N.A. Knott, 2010: Impacts of desalination plant discharges on the marine environment: A critical review of published studies. Water Research, 44(18), 5117-5128. Robertson, M., 2010: Agricultural productivity in Australia and New Zealand: trends, constraints and opportunities. In: Food Security from Sustainable Agriculture. 15th Agronomy Conference, 15-18 November 2010, Lincoln, New Zealand [Dove, H., Culvenor, R.A. (eds.)], Australian Society of Agronomy, Gosford, NSW. Rogan, R., M. O Connor, P. Horwitz, 2005: Nowhere to hide: Awareness and perceptions of environmental change, and their influence on relationships with place. Journal of Environmental Psychology, 25(2), 147-158. Rogers, K., N. Saintilan, C. Copeland, 2012: Modelling wetland surface elevation dynamics and its application to forecasting the effects of sea-level rise on estuarine wetlands. Ecological Modelling, 244, 148-157. Roiko, A., R.B. Mangoyana, S. McFallan, R.W. Carter, J. Oliver, T.F. Smith, 2012: Socio-economic trends and climate change adaptation: the case of South East Queensland. Australasian Journal of Environmental Management, 19(1), 35-50. Rolfe, J., 2009: Climate change and security: the defence component. In: Climate change and security. Planning for the future [Boston, J., Nel, P., Righarts, M. (eds.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, NZ, pp. 93-100. Roman, C.E., A.H. Lynch, D. Dominey-Howes, 2010: Uncovering the Essence of the Climate Change Adaptation Problem A Case Study of the Tourism Sector at Alpine Shire, Victoria, Australia. Tourism Planning & Development, 7(3), 237-252. Ross, H., C. Grant, C.J. Robinson, A. Izurieta, D. Smyth, P. Rist, 2009: Co-management and Indigenous protected areas in Australia: achievements and ways forward. Australasian Journal of Environmental Management, 16(4), 242-252. Rosselló-Nadal, J., A. Riera-Font, V. Cárdenas, 2011: The impact of weather variability on British outbound flows. Climatic Change, 105(1), 281-292. Roura-Pascual, N., C. Hui, T. Ikeda, G. Leday, D.M. Richardson, S. Carpintero, X. Espadaler, C. Gómez, B. Guénard, S. Hartley, P. Krushelnycky, P.J. Lester, M.A. McGeoch, S.B. Menke, J.S. Pedersen, J.P.W. Pitt, J. Reyes, N.J. Sanders, A.V. Suarez, Y. Touyama, D. Ward, P.S. Ward, S.P. Worner, 2011: Relative roles of climatic suitability and anthropogenic influence in determining the pattern of spread in a global invader. Proceedings of the National Academy of Sciences, 108(1), 220-225. Rouse, H.L., N. Norton, 2010: Managing Scientific Uncertainty for Resource Management Planning in New Zealand. Australasian Journal of Environmental Management, 17(2), 66-76. Rouse, H.L., P. Blackett, 2011: Coastal Adaptation to Climate Change. Engaging communities: making it work. NIWA project report CACC125. National Institute of Water and Atmospheric Research (NIWA), Christchurch, 75 pp. Ruddell, A.R., 1995: Recent glacier and climate change in the New Zealand Alps. PhD thesis. School of Earth Sciences, University of Melbourne, Melbourne, 402 pp. Russell, B.D., J.A.I. Thompson, L.J. Falkenberg, S.D. Connell, 2009: Synergistic effects of climate change and local stressors: CO2 and nutrient-driven change in subtidal rocky habitats. Global Change Biology, 15(9), 2153- 2162. Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Russell, S.L., A. Greenaway, F. Carswell, S. Weaver, 2013: Moving beyond mitigation and adaptation : examining climate change responses in New Zealand. Local Environment, advance online, DOI: 10.1080/13549839.2013.792047. Russell-Smith, J., B.P. Murphy, C.P. Meyer, G.D. Cook, S. Maier, A.C. Edwards, J. Schatz, P. Brocklehurst, 2009: Improving estimates of savanna burning emissions for greenhouse accounting in northern Australia: limitations, challenges, applications. International Journal of Wildland Fire, 18(1), 1-18. Rutledge, D.T., R.J. Sinclair, A. Tait, J. Poot, M. Dresser, S. Greenhalgh, M. Cameron, 2011: Triggers and Thresholds of Land-Use Change in Relation to Climate Change and Other Key Trends: A Review and Assessment of Potential Implications for New Zealand. Report prepared for the New Zealand Ministry of Agriculture and Forestry. Landcare Research, Wellington, New Zealand, 130 pp. Sadras, V.O., P.R. Petrie, 2011: Climate shifts in south-eastern Australia: Early maturity of Chardonnay, Shiraz and Cabernet Sauvignon is associated with early onset rather than faster ripening. Australian Journal of Grape and Wine Research, 17(2), 199-205. Salinger, M.J., J.A. Renwick, A.B. Mullan, 2001: Interdecadal Pacific Oscillation and South Pacific climate. International Journal of Climatology, 21(14), 1705-1721. Salinger, M.J., G. Griffiths, A. Gosai, 2005: Extreme pressure differences at 0900 NZST and winds across New Zealand. International Journal of Climatology, 25(9), 1203-1222. Sanders, D., J. Laing, M. Houghton, 2008: Impact of bushfires on tourism and visitation in alpine national parks. Sustainable Tourism Cooperative Research Centre, Gold Coast, Qld, 42 pp. Sato, Y., D.G. Bourne, B.L. Willis, 2009: Dynamics of seasonal outbreaks of black band disease in an assemblage of Montipora species at Pelorus Island (Great Barrier Reef, Australia). Proceedings of the Royal Society B- Biological Sciences, 276(1668), 2795-2803. Sattler, P., M. Taylor, 2008: Building Nature's Safety Net 2008: Progress on the directions for the National Reserve System. WWF-Australia Report. WWF-Australia, Sydney, 64 pp. Saunders, C., W. Kaye-Blake, L. Marshall, S. Greenhalgh, M. de Aragao Pereira, 2009: Impacts of a United States biofuel policy on New Zealand's agricultural sector. Energy Policy, 37(9), 3448-3454. Saunders, C., W. Kaye-Blake, J. Turner, 2010: Modelling Climate Change Impacts on Agriculture and Forestry with the Extended LTEM (Lincoln Trade and Environment Model). Ministry of Agriculture and Forestry, Wellington, NZ, 31 pp. Saunders, D.A., P. Mawson, R. Dawson, 2011: The impact of two extreme weather events and other causes of death on Carnaby's black cockatoo: a promise of things to come for a threatened species? Pacific Conservation Biology, 17, 141-148. SCCCWEA, 2009: Managing our coastal zone in a changing climate: The time to act is now. House of Representatives Standing Committee on Climate Change, Water, Environment and the Arts report on the inquiry into climate change and environmental impacts on coastal communities. Parliament of Australia, Canberra, 368 pp. Scealy, B., D. Newth, D. Gunasekera, J. Finnigan, 2012: Effects of variation in the grains sector response to climate change: an integrated assessment. Economic Papers, 31(3), 327-336. Schiel, D.R., 2011: Biogeographic patterns and long-term changes on New Zealand coastal reefs: Non-trophic cascades from diffuse and local impacts. Journal of Experimental Marine Biology and Ecology, 400(1-2), 33- 51. Schiff, A., S. Becken, 2011: Demand Elasticities for Tourism in New Zealand. Tourism Management, 32(3), 564- 575. Schofield, N., 2011: Climate Change and its Impacts - current understanding, future directions. In: Basin futures: water reform in the Murray-Darling Basin [Grafton, Q., Connell, D. (eds.)]. ANU E-Press, Canberra, pp. 81- 100. Schrobback, P., D. Adamson, J. Quiggin, 2011: Turning Water into Carbon: Carbon Sequestration and Water Flow in the Murray Darling Basin. Environmental and Resource Economics, 49(1), 23-45. Schumann, N., N.J. Gales, R.G. Harcourt, J.P.Y. Arnould, 2013: Impacts of climate change on Australian marine mammals. Australian Journal of Zoology, 61, 146-159. Schuster, S., 2013: Natural hazards and insurance. In: Climate Adaptation Futures [Palutikof, J., Boulter, S., Ash, A., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 133-140. Scott, D., S. Gössling, C.M. Hall, 2012: International tourism and climate change. Wiley Interdisciplinary Reviews: Climate Change, 3(3), 213-232. Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 SCRGSP, 2011: Overcoming Indigenous disadvantage: Key indicators 2011. Steering Committee for the Review of Government Service Provision, Productivity Commission, Canberra, 64 pp. Scutella, R., R. Wilkins, W. Kostenko, 2009: Estimates of Poverty and Social Exclusion in Australia: A Multidimensional Approach. Melbourne Institute of Applied Economic and Social Research Working Paper No. 26/09. Melbourne University, Melbourne, 67 pp. Serrao-Neumann, S., F. Crick, B. Harman, M. Sano, O. Sahin, R. Staden, G. Schuch, S. Baum, D. Low Choy, 2013: Improving cross-sectoral climate change adaptation for coastal settlements: insights from South East Queensland, Australia. Regional Environmental Change, advance online, DOI: 10.1007/s10113-013-0442-6. SGS, 2012: Tasmanian Coastal Adaptation Pathways Project. Lauderdale Recommended Actions. Report prepared by SGS Economics & Planning. Local Government Association of Tasmania, Hobart, 55 pp. Shakesby, R.A., P.J. Wallbrink, S.H. Doerr, P.M. English, C.J. Chafer, G.S. Humphreys, W.H. Blake, K.M. Tomkins, 2007: Distinctiveness of wildfire effects on soil erosion in south-east Australian eucalypt forests assessed in a global context. Forest Ecology and Management, 238(1-3), 347-364. ShapeNZ, 2009: New Zealanders attitudes to climate change. New Zealand Business Council for Sustainable Development, Auckland, 34 pp. Sharma, V., S.v.d. Graaff, B. Loechel, D. Franks, 2013: Extractive resource development in a changing climate. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 116 pp. Shearer, H., 2011: Using Geographic Information Systems to explore the determinants of household water consumption and response to the Queensland Government demand-side policy measures imposed during the drought of 2006-2008. In: State of Australian Cities National Conference, 29 November-2 December 2011, Melbourne, Vic [Whitzman, C., Fincher, R. (eds.)], Faculty of the Built Environment, University of New South Wales, Sydney, 15 pp. Sheppard, C.S., 2012: Potential spread of recently naturalised plants in New Zealand under climate change. Climatic Change, advance online, doi: 10.1007/s10584-012-0605-3. Shoo, L.P., C. Storlie, J. Vanderwal, J. Little, S.E. Williams, 2011: Targeted protection and restoration to conserve tropical biodiversity in a warming world. Global Change Biology, 17(1), 186-193. Simioni, G., P. Ritson, M.U.F. Kirschbaum, J. McGrath, I. Dumbrell, B. Copeland, 2009: The carbon budget of Pinus radiata plantations in south-western Australia under four climate change scenarios. Tree Physiology, 29(9), 1081-1093. Sinclair, E., 2008: The Changing Climate of New Zealand s Security: Risk and Resilience in a Climate Affected Security Environment. Working Paper 08/11. Institute of Policy Studies, Victoria University, Wellington, NZ, 68 pp. Singh, S., S. Davey, M. Cole, 2010: Implications of climate change for forests, vegetation and carbon in Australia. New Zealand Journal of Forestry Science, 40, 141-152. Skinner, R., 2010: Adaptation to Climate Change in Melbourne: Changing the Fundamental Planning Assumptions. Presented at International Adaptation Forum on Climate Change Impacts on Water Supply, Washington DC, January, 2010. Melbourne Water, Melbourne, 24 pp. Smale, D.A., T. Wernberg, 2013: Extreme climatic event drives range contraction of a habitat-forming species. Proceedings of the Royal Society of London, Series B: Biological Sciences, 280(1754), 20122829. Smales, I., 2006: Wind farm collision risk for birds. Cumulative risks for threatened and migratory species. Report by Biosis Research Pty Ltd. Department of Environment and Heritage, Canberra, 237 pp. Smart, A., A. Aspinall, 2009: Water and the electricity generation industry. Waterlines report No. 18. National Water Commission, Canberra, 115 pp. Smart, R.E., 2010: A lump of coal, a bunch of grapes. Journal of Wine Research, 21(2), 107-111. Smart, G.M., A.I. McKerchar, 2010: More flood disasters in New Zealand. Journal of Hydrology (New Zealand), 49(2), 69-78. Smith, T.F., C. Brooke, T.G. Measham, B. Preston, R. Gorddard, G. Withycombe, B. Beveridge, C. Morrison, 2008: Case studies of adaptive capacity: systems approach to regional climate change adaptation strategies. Sydney Coastal Councils Group Inc, Sydney, NSW. Smith, I.N., B. Timbal, 2010: Links between tropical indices and southern Australian rainfall. International Journal of Climatology, 32(1), 33-40. Smith, T., T. Lynam, B. Preston, J. Matthews, R.W. Carter, D. Thomsen, J. Carter, A. Roiko, R. Simpson, P. Waterman, M. Bussey, N. Keys, C. Stephenson, 2010: Towards Enhancing Adaptive Capacity for Climate Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Change Response in South East Queensland. The Australasian Journal of Disaster and Trauma Studies, 2010-1, online. Smith, H.G., G.J. Sheridan, P.N.J. Lane, P. Nyman, S. Haydon, 2011a: Wildfire effects on water quality in forest catchments: A review with implications for water supply. Journal of Hydrology, 396(1 2), 170-192. Smith, W., C. Davies-Colley, A. Mackay, G. Bankoff, 2011b: Social impact of the 2004 Manawatu floods and the hollowing out of rural New Zealand. Disasters, 35(3), 540 553. Smith, T., D. Low Choy, D.C. Thomsen, S. Neumann, F. Crick, M. Sano, R. Richards, B. Harman, S. Baum, S. Myers, V. Sharma, M. Bussey, J. Matthews, A. Roiko, R.W. Carter, 2014: Adapting Australian coastal regions to climate change: A case study of South East Queensland. In: Climate Change and the Coastal Zone: Building Resilient Communities [Glavovic, B., Kay, R.C., Kelly, M., Travers, A. (eds.)]. CRC Press, London, www.crcpress.com/product/isbn/9780415464871. SNZ, 2010a: National ethnic population projections: 2006 (base) - 2026 update. Statistics New Zealand, Wellington, 31 pp. SNZ, 2010b: Are New Zealanders living closer to the coast? Internal Migration. Statistics New Zealand, Wellington, 4 pp. SNZ, 2010c: Subnational Population Projections: 2006 (base) 2031 update. Statistics New Zealand, Wellington, 15 pp. SNZ, 2011: National Accounts: Long-term data series. Table E.1.2 real gross domestic product (Excel data spreadsheet, online at www.stats.govt.nz). Statistics New Zealand, Wellington. SNZ, 2012a: National Population Projections: 2011 (base) 2061. Statistics New Zealand, Wellington, 28 pp. SNZ, 2012b: Global New Zealand. International Trade, Investment, and Travel Profile: year ended June 2012. Ministry of Foreign Affairs and Trade and Statistics New Zealand, Wellington, NZ, 199 pp. SNZ, 2013: New Zealand's seafood industry - Infographic. Statistics New Zealand, Wellington, 1 pp. SoE, 2011: State of the Environment 2011. Independent report to the Australian Government Minister for Sustainability, Environment, Water, Population and Communities. Australian State of the Environment Committee, Canberra, 932 pp. Soste, L., 2010: Victorian Climate Change Adaptation Program (VCCAP): final report - scenario theme. Department of Primary Industries, State Government of Victoria, Tatura, Vic. Spickett, J., H. Brown, D. Katscherian, 2008: Health impacts of climate change: Adaptation strategies for Western Australia. Western Australia Department of Health, Perth, 74 pp. Spillman, C.M., 2011: Operational real-time seasonal forecasts for coral reef management. Journal of Operational Oceanography, 4(1), 13-22. Srinivasan, M.S., J. Schmidt, S. Poyck, E. Hreinsson, 2011: Irrigation Reliability Under Climate Change Scenarios: A Modeling Investigation in a River-Based Irrigation Scheme in New Zealand. Journal of the American Water Resources Association, 47(6), 1261-1274. Stafford-Smith, M., L. Horrocks, A. Harvey, C. Hamilton, 2011: Rethinking adaptation for a 4°C world. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1934), 196-216. Stafford-Smith, M., 2013: Scenarios for picturing a future adapted to +4°C. In: Climate Adaptation Futures [Palutikof, J., Boulter, S.L., Ash, A.J., Stafford-Smith, M., Parry, M., Waschka, M., Guitart, D. (eds.)]. John Wiley & Sons, Brisbane, pp. 119-125. Stamatov, V., A. Stamatov, 2010: Long-term impact of water desalination plants on the energy and carbon dioxide balance of Victoria, Australia: a case study from Wonthaggi. Water and Environment Journal, 24(3), 208-214. Standards Australia, 2013: Climate change adaptation for settlements and infrastructure - A risk based approach. AS 5334-2013. Standards Australia, Canberra, 47 pp. Stark, C., K. Penney, A. Feng, 2012: 2012 Australian Energy Update. Bureau of Resources and Energy Economics, Canberra, 26 pp. Steffen, W., A.A. Burbidge, L. Hughes, R. Kitching, D. Lindenmayer, W. Musgrave, M. Stafford-Smith, P.A. Werner, 2009: Australia's biodiversity and climate change: A strategic assessment of the vulnerability of Australia's biodiversity to climate change. A report to the Natural Resource Management Ministerial Council commissioned by the Australian Government. CSIRO Publishing, Collingwood, Vic, 289 pp. Steffen, W., J. Sims, J. Walcott, G. Laughlin, 2011: Australian agriculture: Coping with dangerous climate change. Regional Environmental Change, 11(Suppl. 1), 205-214. Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Stehlik, D., G. Lawrence, I. Gray, 2000: Gender and Drought: Experiences of Australian Women in the Drought of the 1990s. Disaster, 24(1), 38-53. Stephenson, J., B. Barton, G. Carrington, D. Gnoth, R. Lawson, P. Thorsnes, 2010: Energy cultures: A framework for understanding energy behaviours. Energy Policy, 38(10), 6120-6129. Stewart, H.T.L., D.H. Race, A.L. Curtis, 2011: New Forests in Changing Landscapes in South-East Australia. International Forestry Review, 13(1), 67-79. Stewart, M.G., X. Wang, 2011: Risk Assessment of Climate Adaptation Strategies for Extreme Wind Events in Queensland. CSIRO, Canberra, 85 pp. Stewart, M.G., X. Wang, M.N. Nguyen, 2012: Climate change adaptation for corrosion control of concrete infrastructure. Structural Safety, 35(0), 29-39. Stokes, C., A. Ash, 2007: Impacts of climate change on marginal tropical animal production systems. In: Agroecosystems in a Changing Climate [Newton, P.C.D., Carran, R.A., Edwards, G.R., Niklaus, P.A. (eds.)]. CRC Press, New York, pp. 323-328. Stork, N.E., J. Balston, G.D. Farquhar, P.J. Franks, J.A.M. Holtum, M.J. Liddell, 2007: Tropical rainforest canopies and climate change. Austral Ecology, 32(1), 105-112. Strengers, Y., 2008: Comfort expectations: the impact of demand-management strategies in Australia. Building Research and Information, 36(4), 381-391. Strengers, Y., C. Maller, 2011: Integrating health, housing and energy policies: Social practices of cooling. Building Research and Information, 39(Compendex), 154-168. Stroombergen, A., A. Tait, J. Renwick, K. Patterson, 2006: The relationship between New Zealand's climate, energy, and the economy in 2025. Kotuitui New Zealand Journal of Social Sciences Online, 1, 139-160. Stroombergen, A., 2010: The International Effects of Climate Change on Agricultural Commodity Prices, and the Wider Effects on New Zealand. Working Paper 10-14. Motu Economic and Public Policy Research, Wellington, New Zealand, 37 pp. Suncorp, 2013: Risky Business. Insurance and Natural Disaster Risk Management. Suncorp Personal Insurance Public Policy. Suncorp, Brisbane, Qld, 20 pp. Sutherst, R.W., B.S. Collyer, T. Yonow, 2000: The vulnerability of Australian horticulture to the Queensland fruit fly, Bactrocera (Dacus) tryoni, under climate change. Australian Journal of Agricultural Research, 51(4), 467- 480. Sutton, P.J.H., M. Bowen, D. Roemmich, 2005: Decadal temperature changes in the Tasman sea. New Zealand Journal of Marine and Freshwater Research, 39(6), 1321-1329. Syed, A., 2012: Australian Energy Projections to 2050. Bureau of Resources and Energy Economics, Canberra, 66 pp. Tait, A., 2008: Future projections of growing degree days and frost in New Zealand and some implications for grape growing. Weather and Climate, 28, 17-36. Tait, A., T. Baisden, D. Wratt, B. Mullan, A. Stroombergen, 2008a: An initial assessment of the potential effects of climate change on New Zealand agriculture. New Zealand Science Review, 65(3), 50-56. Tait, A., J. Sturman, B. Mullan, D. King, G. Griffiths, P. Newsome, G. Harmsworth, I. Nicholas, L. Gea, N. Porter, J. Reid, 2008b: Use of climate, soil, and tree species information for identifying land use options across the Gisborne district. National Institute of Water and Atmospheric Research (NIWA), Wellington, 104 pp. Taptiklis, N., 2011: Climate resilient water management in Wellington, New Zealand. NZCCRI 2011 Report 09. New Zealand Climate Change Research Institute, Victoria University of Wellington, Wellington, 62 pp. Taylor, M.A.P., M. Philp, 2010: Adapting to climate change - implications for transport infrastructure, transport systems and travel behaviour. Road & Transport Research, 19(4), 66-79. Taylor, R., D. Rutledge, H. van Roon, 2011: Building capacity in urban sustainability assessment through use of a scenarios game. Journal of Education for Sustainable Development, 5(1), 75-87. Taylor, B.M., B.P. Harman, S. Heyenga, R.R.J. McAllister, 2012a: Property Developers and Urban Adaptation: Conceptual and Empirical Perspectives on Governance. Urban Policy and Research, 30(1), 5-24. Taylor, S., L. Kumar, N. Reid, 2012b: Impacts of climate change and land-use on the potential distribution of an invasive weed: A case study of Lantana camara in Australia. Weed Research, 52(5), 391-401. Te Aho, L., 2007: Contemporary issues in M ori law and society. Waikato Law Review, 15, 138-158. Teixeira, E.I., H.E. Brown, A. Fletcher, G. Hernandez-Ramirez, A. Soltani, S. Viljanen-Rollinson, A. Horrocks, P. Johnstone, 2012: Broadacre cropping: Adapting broad acre cropping to climate change. In: Enhanced climate Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 change impact and adaptation evaluation: A comprehensive analysis of New Zealand's land-based primary sectors. [Clark, A., Nottage, R. (eds.)]. Ministry for Primary Industries, New Zealand, pp. 189-236. Teng, J., F. Chiew, J. Vaze, S. Marvanek, D.G.C. Kirono, 2012: Estimation of climate change impact on mean annual runoff across continental Australia using Budyko and Fu equations and comparison with hydrological model simulations. Journal of Hydrometeorology, advance online, doi:10.1175/JHM-D-11-097.1. TGA, 2009: The Geneva Reports - Risk and Insurance Reseach. No. 2. The insurance industry and climate change Contribution to the global debate. The Geneva Association, Geneva, Switzerland, 152 pp. Thatcher, M.J., 2007: Modelling changes to electricity demand load duration curves as a consequence of predicted climate change for Australia. Energy, 32(Compendex), 1647-1659. Thomas, D.T., J. Sanderman, S.J. Eady, D.G. Masters, P. Sanford, 2012: Whole Farm Net Greenhouse Gas Abatement from Establishing Kikuyu-Based Perennial Pastures in South-Western Australia. Animals, 2(3), 316- 330. Thompson, D., J. Wallace, 2000: Annular modes in the extratropical circulation. Part I: Month-to-month variability. Journal of Climate, 13(5), 1000-1016. Thompson, P.A., M.E. Baird, T. Ingleton, M.A. Doblin, 2009: Long-term changes in temperate Australian coastal waters: implications for phytoplankton. Marine Ecology-Progress Series, 394, 1-19. Thomson, L.J., S. Macfadyen, A.A. Hoffmann, 2010: Predicting the effects of climate change on natural enemies of agricultural pests. Biological Control, 52(3), 296-306. Thorburn, P.J., M.J. Robertson, B.E. Clothier, V.O. Snow, E. Charmley, J. Sanderman, E. Teixeira, R.A. Dynes, A. Hall, H. Brown, S.M. Howden, M. Battaglia, 2012: Australia and New Zealand Perspectives on Climate Change and Agriculture. In: Agriculture's Contributions to Climate Change Solutions: Mitigation and Adaptation at Global and Regional Scales. Handbook of Climate Change and Agroecosystems, Vol 2 [Rosenzweig, C., Hillel, D. (eds.)]. American Society of Agronomy and Imperial College Press, New York, pp. 107-141. Thresher, R.E., J.A. Koslow, A.K. Morison, D.C. Smith, 2007: Depth-mediated reversal of the effects of climate change on long-term growth rates of exploited marine fish. Proceedings of the National Academy of Sciences of the United States of America, 104(18), 7461-7465. Timbal, B., J.M. Arblaster, S. Power, 2006: Attribution of the late-twentieth-century rainfall decline in southwest Australia. Journal of Climate, 19(10), 2046-2062. Timbal, B., D.A. Jones, 2008: Future projections of winter rainfall in southeast Australia using a statistical downscaling technique. Climatic Change, 86(1-2), 165-187. Timbal, B., J. Arblaster, K. Braganza, E. Fernandez, H. Hendon, B. Murphy, M. Raupach, C. Rakich, I. Smith, K. Whan, M. Wheeler, 2010a: Understanding the anthropogenic nature of the observed rainfall decline across south eastern Australia. CAWCR Technical Report No. 26. Centre for Australian Weather and Climate Research, Melbourne, Vic, 202 pp. Timbal, B., R. Kounkou, G.A. Mills, 2010b: Changes in the Risk of Cool-Season Tornadoes over Southern Australia due to Model Projections of Anthropogenic Warming. Journal of Climate, 23(9), 2440-2449. Todd, M., W. Zhang, S. Kerr, 2009: Competition for land between biofuels, pastoral agriculture. In: Bioenergy options for New Zealand. Analysis of large-scale bioenergy from forestry [Hall, P., Jack, M. (eds.)]. Scion, Rotorua, NZ, pp. 122-140. Tolhurst, K., 2009: Report of the Physical Nature of the Victorian Fires Occurring on 7 February 2009. Report for the Counsel Assisting the Royal Commission, Victorian Bushfire Royal Commission, Melbourne, 18 pp. Tong, S.L., C.Z. Ren, N. Becker, 2010a: Excess deaths during the 2004 heatwave in Brisbane, Australia. International Journal of Biometeorology, 54(4), 393-400. Tong, S.L., X.Y. Wang, A.G. Barnett, 2010b: Assessment of Heat-Related Health Impacts in Brisbane, Australia: Comparison of Different Heatwave Definitions. Plos One, 5(8), e12155. Tourism Queensland, 2007: Regional Tourism Crisis Management Plan Template. A guide to preparing a Regional Tourism Crisis Management Plan. Tourism Queensland, Brisbane, Qld, 30 pp. Tourism Queensland, 2010: Climate Futures Industry Tool. Tourism Queensland, Brisbane, Qld, 31 pp. Tourism Victoria, 2010: Crisis Essentials. Crisis Management for tourism businesses. Tourism Victoria, Melbourne, Vic, 18 pp. TPK, 2007: A time for change in M ori economic development. Te Puni Kokiri, Wellington, New Zealand, 36 pp. TRA, 2010: Impact of the drought on tourism in the Murray River region: Summary of results. Tourism Research Australia, Department of Resources, Energy and Tourism, Canberra, 3 pp. Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Traill, L.W., K. Perhans, C.E. Lovelock, A. Prohaska, S. McFallan, J.R. Rhodes, K.A. Wilson, 2011: Managing for change: wetland transitions under sea-level rise and outcomes for threatened species. Diversity and Distributions, 17(6), 1225-1233. Trewin, B., H. Vermont, 2010: Changes in the frequency of record temperatures in Australia, 1957-2009. Australian Meteorological and Oceanographic Journal, 60(2), 113-120. Troccoli, A., K. Muller, P. Coppin, R. Davy, C. Russell, A.L. Hirsch, 2012: Long-term wind speed trends over Australia. Journal of Climate, 25, 170-183. Trotman, R., 2008: Project Twin Streams to 2007. Waitakere City Council's story. Waitakere City Council, Waitakere, 56 pp. Troy, P. (eds.), 2008: Troubled waters: confronting the water crisis in Australia s cities. ANU E-Press, Canberra, 217 pp. Tryhorn, L., J. Risbey, 2006: On the distribution of heat waves over the Australian region. Australian Meteorological Magazine, 55(3), 169-182. TSRA, 2010: Torres Strait climate change strategy 2010 2013. Environmental Management Program, Torres Strait Regional Authority, Darwin, 20 pp. Turton, S., W. Hadwen, R. Wilson, B. Jorgensen, D. Simmons, 2009: The impacts of climate change on Australian Tourist Destinations. Developing Adaptation and Response Strategies. Sustainable Tourism Cooperative Research Centre, Gold Coast, Qld, 77 pp. Turton, S., T. Dickson, W. Hadwen, B. Jorgensen, T. Pham, D. Simmons, P. Tremblay, R. Wilson, 2010: Developing an approach for tourism climate change assessment: evidence from four contrasting Australian case studies. Journal of Sustainable Tourism, 18(3), 429-448. UN, 2012: World Urbanization Prospects. The 2011 Revision. Department of Economic and Social Affairs, United Nations Secretariat, New York, 318 pp. Valenzuela, E., K. Anderson, 2011: Climate change and food security to 2030: A global economy-wide perspective. Economia Agraria y Recursos Naturales, 11(1), 29-58. van den Honert, R.C., J. McAneney, 2011: The 2011 Brisbane Floods causes, impacts and implications. Water, 3, 1149-1173. van der Wal, J., H.T. Murphy, A.S. Kutt, G.C. Perkins, B.L. Bateman, J.J. Perry, A.R. Reside, 2013: Focus on poleward shifts in species' distribution underestimates the fingerprint of climate change. Nature Climate Change, 3, 239-243. van Dijk, A., R. Evans, P. Hairsine, S. Khan, R. Nathan, Z. Paydar, N. Viney, L. Zhang, 2006: Risks to shared water resources of the Murray-Darling Basin. Publication 22/06. Murray Darling Basin Commission, Canberra, 49 pp. Vaneckova, P., P.J. Beggs, R.J. de Dear, K.W.J. McCracken, 2008: Effect of temperature on mortality during the six warmer months in Sydney, Australia, between 1993 and 2004. Environmental Research, 108(3), 361-369. VBRC, 2010: 2009 Victorian Bushfires Royal Commission Final Report - Summary. 2009 Victorian Bushfires Royal Commission, Melbourne, 42 pp. Veland, S., R. Howitt, D. Dominey-Howes, 2010: Invisible institutions in emergencies: Evacuating the remote Indigenous community of Warruwi, Northern Territory Australia, from Cyclone Monica. Environmental Hazards-Human and Policy Dimensions, 9(2), 197-214. Verdon, D.C., A.M. Wyatt, A.S. Kiem, S.W. Franks, 2004: Multidecadal variability of rainfall and streamflow: Eastern Australia. Water Resources Research, 40(10), W10201. Verdon-Kidd, D.C., A.S. Kiem, E.K. Austin, 2012: Decision making under uncertainty - Bridging the gap between end user needs and climate science capability. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 116 pp. Veron, J.E.N., O. Hoegh-Guldberg, T.M. Lenton, J.M. Lough, D.O. Obura, P. Pearce-Kelly, C.R.C. Sheppard, M. Spalding, M.G. Stafford-Smith, A.D. Rogers, 2009: The coral reef crisis: The critical importance of <350 ppm CO2. Marine Pollution Bulletin, 58(10), 1428-1436. Verschuuren, J., J. McDonald, 2012: Towards a legal framework for coastal adaptation: assessing the first steps in Europe and Australia. Transnational Environmental Law, 1(2), 355-379. Victorian Government, 2009a: January 2009 Heatwave in Victoria: An Assessment of Health Impacts. Department of Human Services, Melbourne, Vic, 16 pp. Victorian Government, 2009b: Heatwave Plan for Victoria 2009-2010: Protecting health and reducing harm from heatwaves. Department of Health, Melbourne, Vic, 42 pp. Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Walker, S.J., T.A. Schlacher, L.M.C. Thompson, 2008: Habitat modification in a dynamic environment: The influence of a small artificial groyne on macrofaunal assemblages of a sandy beach. Estuarine, Coastal and Shelf Science, 79(1), 24-34. Wan, S., R.J. Norby, J. Ledford, J.F. Weltzin, 2007: Responses of soil respiration to elevated CO2, air warming, and changing soil water availability in an old-field grassland. Global Change Biology, 13(11), 2411-2424. Wang, X., D. Chen, Z.G. Ren, 2010a: Assessment of climate change impact on residential building heating and cooling energy requirement in Australia. Building and Environment, 45(7), 1663-1682. Wang, X., M. Stafford-Smith, R. McAllister, A. Leitch, S. McFallan, S. Meharg, 2010b: Coastal inundation under climate change: a case study in South East Queensland. Report prepared for the South East Queensland Climate Adaptation Research Initiative. Climate Adaptation Flagship Working Paper No. 6. CSIRO, Brisbane, 28 pp. Wang, X., R. McAllister, 2011: Adapting to Heatwaves and Coastal Flooding. In: Climate Change: Science and Solutions for Australia [Cleugh, H., Stafford-Smith, M., Battaglia, M., Graham, P. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 73-84. Wang, X., M. Stewart, M. Nguyen, M. Syme, 2012: Implications of Climate Change to the Durability of Concrete Structures - An Investigation by Simulations. Concrete in Australia, 36(4), 27-35. Watanabe, T., S. Bowatte, P.C.D. Newton, 2013: Reduced nifH gene and transcript numbers indicate a reduced capacity for nitrogen fixation in nodules of white clover after exposure of a grazed grassland to elevated CO2. Biogeosciences Discussion, 10, 9867-9896. Watson, R.A., G.B. Nowara, S.R. Tracey, E.A. Fulton, C.M. Bulman, G.J. Edgar, N.S. Barrett, J.M. Lyle, S.D. Frusher, C.D. Buxton, 2012: Ecosystem model of Tasmanian waters explores impacts of climate change induced changes in primary productivity. Ecological Modelling, advance online, doi: 10.1016/j.ecolmodel.2012.05.008. Watt, M.S., M.U.F. Kirschbaum, T.S.H. Paul, A. Tait, H.G. Pearce, E.G. Brockerhoff, J.R. Moore, L.S. Bulman, D.J. Kriticos, 2008: The effect of climate change on New Zealand's planted forests: impacts, risks and opportunities. Scion Client Report CC MAF POL 2008-07 (106-1) No. 1. Scion, Rotorua, 149 pp. Watt, M.S., D.J. Palmer, L.S. Bulman, 2011a: Predicting the severity of Dothistroma needle blight on Pinus radiata under climate change. New Zealand Journal of Forestry Science, 41, 207-215. Watt, M.S., J.K. Stone, I.A. Hood, L.K. Manning, 2011b: Using a climatic niche model to predict the direct and indirect impacts of climate change on the distribution of Douglas-fir in New Zealand. Global Change Biology, 17(12), 3608-3619. Watterson, I.G., 2012: Understanding and partioning future climates for Australian regions from CMIP3 using ocean warming indices. Climatic Change, 111, 903-922. WCC, 2010: Wellington City s 2010 climate change action plan. Wellington City Council, Wellington, 39 pp. Webb, L.B., P.H. Whetton, E.W.R. Barlow, 2008: Climate change and winegrape quality in Australia. Climate Research, 36(2), 99-111. Webb, L.B., P.H. Whetton, J. Bhend, R. Darbyshire, P.R. Briggs, E.W.R. Barlow, 2012a: Earlier wine-grape ripening driven by climatic warming and drying and management practices. Nature Climate Change, 2, 259 264. Webb, N.P., C.J. Stokes, J.C. Scanlan, 2012b: Interacting effects of vegetation, soils and management on the sensitivity of Australian savanna rangelands to climate change. Climatic Change, 112(3-4), 925-943. Webb, B., J.-L. Beh, 2013: Leading adaptation practices and support strategies for Australia: An international and Australian review of products and tools. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, Qld, 84 pp. Webb, B., R. McKellar, R. Kay, 2013: Climate change adaptation in Australia: experience, challenges and capacity- building. Australian Journal of Emergency Management, advance online, DOI: 10.1080/14486563.2013.835285. Weber, E.P., A. Memon, B. Painter, 2011: Science, Society, and Water Resources in New Zealand: Recognizing and Overcoming a Societal Impasse. Journal of Environmental Policy & Planning, 13(1), 49-69. Webster, T., J. Morison, N. Abel, E. Clark, L. Rippin, A. Herr, B. Taylor, P. Stone, 2009: Irrigated agriculture: development opportunities and implications for northern Australia. In: Northern Australia Land and Water Science Review 2009. Report to the Northern Territories Land and Water Task Force [CSIRO Sustainable Agriculture Flagship (eds.)]. Department of Infrastructure, Transport, Regional Development and Local Government, Canberra, pp. 10-1 to 10-53. Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Welbergen, J.A., S.M. Klose, N. Markus, P. Eby, 2008: Climate change and the effects of temperature extremes on Australian flying-foxes. Proceedings of the Royal Society B-Biological Sciences, 275(1633), 419-425. Wernberg, T., B.D. Russell, M.S. Thomsen, C.F.D. Gurgel, C.J.A. Bradshaw, E.S. Poloczanska, S.D. Connell, 2011: Seaweed Communities in Retreat from Ocean Warming. Current Biology, 21(21), 1828-1832. Wernberg, T., D.A. Smale, F. Tuya, M.S. Thomsen, T.J. Langlois, T. de Bettignies, S. Bennett, C.S. Rousseaux, 2013: An extreme climatic event alters marine ecosystem structure in a global biodiversity hotspot. Nature Climate Change, 3, 78-82. Westra, S., S. Sisson, 2011: Detection of non-stationarity in precipitation extremes using a max-stable process model. Journal of Hydrology, 406(1-2), 119-128. Westra, S., 2012: Adapting to Climate Change Revising our Approach to Estimating Future Floods. University of New South Wales, Sydney, 25 pp. Westra, S., J. Evans, R. Mehrotra, A. Sharma, 2013: A conditional disaggregation algorithm for generating fine time-scale rainfall data in a warmer climate. Journal of Hydrology, 479, 86-99. White, N.E., 2009: Local Government Planning Responses to the Physical Impacts of Climate Change in New South Wales, Australia. International Journal of Climate Change Impacts and Responses, 1(2), 1-16. White, D.A., D.S. Crombie, J. Kinal, M. Battaglia, J.F. McGrath, D.S. Mendham, S.N. Walker, 2009: Managing productivity and drought risk in Eucalyptus globulus plantations in south-western Australia. Forest Ecology and Management, 259, 33-44. White, N.E., J. Buultjens, 2012: Climate change policy responses of Australia and New Zealand national governments: implications for sustainable tourism. In: Tourism, Climate Change and Sustainability [Reddy, M.V., Wilkes, K. (eds.)]. Earthscan Publications, Oxford, pp. 117-133. Whitehead, P.J., P. Purdon, P.M. Cooke, J. Russell-Smith, S. Sutton, 2009: The West Arnhem Land Fire Abatement (WALFA) project: the institutional environment and its implications. In: Culture, ecology and economy of fire management in north Australia savannas: Rekindling the wurrk tradition [Russell-Smith, J., Whitehead, P.J., Cooke, P.M. (eds.)]. CSIRO Publishing, Collingwood, Vic, pp. 287-312. Whittaker, J., J. Handmer, D. Karoly, 2013: After 'Black Saturday': adapting to bushfires in a changing climate. In: Natural Disasters and Adaptation to Climate Change [Boulter, S., Palutikof, J., Karoly, D., Guitart, D. (eds.)]. Cambridge University Press, Cambridge, UK, pp. 115-136. Wilby, R.L., R.J. Keenan, 2012: Adapting to flood risk under climate change. Progress in Physical Geography, 36(3), 348-378. Wilkinson, S.N., P.J. Wallbrink, G.J. Hancock, W.H. Blake, R.A. Shakesby, S.H. Doerr, 2009: Fallout radionuclide tracers identify a switch in sediment sources and transport-limited sediment yield following wildfire in a eucalypt forest. Geomorphology, 110(3-4), 140-151. Williams, R.J., R.A. Bradstock, G.J. Cary, N.J. Enright, A. Gilll, M,, A.C. Leidloff, C. Lucas, R. Whelan, A.N. Andersen, D.M.J.S. Bowman, P.J. Clarke, G.D. Cook, K. Hennessy, A. York, 2009: Interactions between climate change, fire regimes and biodiversity in Australia - a preliminary assessment. Department of Climate Change, Canberra, 196 pp. Williams, S.E., L.P. Shoo, J.L. Isaac, A.A. Hoffmann, G. Langham, 2008: Towards an Integrated Framework for Assessing the Vulnerability of Species to Climate Change. Plos Biology, 6(12), 2621-2626. Williams, N.S.G., J.P. Rayner, K.J. Raynor, 2010: Green roofs for a wide brown land: Opportunities and barriers for rooftop greening in Australia. Urban Forestry & Urban Greening, 9(3), 245-251. Willis, T.J., S.J. Handley, F.H. Chang, C.S. Law, D.J. Morrisey, A.B. Mullan, M. Pinkerton, K.L. Rodgers, P.J.H. Sutton, A. Tait, 2007: Climate change and the New Zealand marine environment. NIWA Client Report NEL2007-025. Department of Conservation, Wellington, 81 pp. Willsman, A., T. Chinn, J. Hendrikx, A. Lorrey, 2010: New Zealand Glacier Monitoring: End of Summer Snowline Survey 2010. NIWA Client Report no. CHC2010-113. National Institute of Water and Atmospheric Research (NIWA), Christchurch, 132 pp. Wilson, N., 2011: End-of-term review of the New Zealand Government s response to climate change: a public health perspective. New Zealand Medical Journal, 124(1345), 90-95. Wilson, J., S. Becken, 2011: Perceived Deficiencies in the Provision of Climate and Weather Information for Tourism: A New Zealand Media Analysis. New Zealand Geographer, 67(3), 148-160. Wilson, R., S. Turton, 2011a: Climate change adaptation options, tools and vulnerability. Contribution of Work Package 4 to the Forest Vulnerability Assessment. National Climate Change Adaptation Research Facility (NCCARF), Gold Coast, 124 pp. Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wilson, R., S. Turton, 2011b: The impact of climate change on reef-based tourism in Cairns, Australia Adaptation and Response Strategies for a highly vulnerable destination. In: Disappearing Destinations [Jones, A., Phillips, M. (eds.)]. CAB International, Wallingford, pp. 233-253. Winterbourn, M.J., S. Cadbury, C. Ilg, A.M. Milner, 2008: Mayfly production in a New Zealand glacial stream and the potential effect of climate change. Hydrobiologia, 603, 211-219. Wittwer, G., M. Griffith, 2011: Modelling drought and recovery in the southern Murray-Darling basin. Australian Journal of Agricultural and Resource Economics, 55(3), 342-359. Woods, R., A. Tait, B. Mullan, J. Hendrikx, J. Diettrich, 2008: Projected climate and river flow for the Rangitata catchment for 2040. NIWA Client Report: CHC2008-097. National Institute of Water and Atmospheric Research (NIWA), Christchurch, 28 pp. Woodward, A., S. Hales, N. de Wet, 2001: Climate change: potential effects on human health in New Zealand. A report prepared for the Ministry for the Environment as part of the New Zealand Climate Change Programme. Ministry for the Environment, Wellington, New Zealand, 27 pp. Woodward, E., S. Jackson, M. Finn, P.M. McTaggart, 2012: Utilising Indigenous seasonal knowledge to understand aquatic resource use and inform water resource management in northern Australia. Ecological Management & Restoration, 13(1), 58-64. Wooldridge, S.A., T.J. Done, C.R. Thomas, I.I. Gordon, P.A. Marshall, R.N. Jones, 2012: Safeguarding coastal coral communities on the central Great Barrier Reef (Australia) against climate change: realizable local and global actions Climatic Change, 112(3-4), 945-961. Wratt, D., A.B. Mullan, A. Tait, R. Woods, T. Baisden, D. Giltrap, J. Hendy, A. Stroombergen, 2008: Costs and benefits of climate change and adaptation to climate change in New Zealand agriculture: what do we know so far? Contract Report by the Ecoclimate Consortium. Ministry of Agriculture and Forestry, Wellington, 112 pp. Wyborn, C., 2009: Managing change or changing management: climate change and human use in Kosciuszko National Park. Australasian Journal of Environmental Management, 16(4), 208-217. Yates, C.J., J. Elith, A.M. Latimer, D. Le Maitre, G.F. Midgley, F.M. Schurr, A.G. West, 2010a: Projecting climate change impacts on species distributions in megadiverse South African Cape and Southwest Australian Floristic Regions: Opportunities and challenges. Austral Ecology, 35(4), 374-391. Yates, C.J., A. McNeill, J. Elith, G.F. Midgley, 2010b: Assessing the impacts of climate change and land transformation on Banksia in the South West Australian Floristic Region. Diversity and Distributions, 16(1), 187-201. Yeates, G., P.C.D. Newton, 2009: Long-term changes in topsoil nematode populations in grazed pasture under elevated atmospheric carbon dioxide. Biology and Fertility of Soils, 45(8), 799-808. Yuen, E., S. Jovicich, B. Preston, 2012: Climate change vulnerability assessments as catalysts for social learning: four case studies in south-eastern Australia. Mitigation and Adaptation Strategies for Global Change, advance online, doi:10.1007/s11027-012-9376-4. Zammit, C., R. Woods, 2011: Projected climate and river flow for the Ashley catchment for 2040 and 2090. NIWA Client Report: CHC2010-160. National Institute of Water and Atmospheric Research (NIWA), Christchurch, 46 pp. Zander, K.K., L. Petheram, S.T. Garnett, 2013: Stay or leave? Potential climate change adaptation strategies among Aboriginal people in coastal communities in northern Australia. Natural Hazards, 67(2), 591-609. Zeppel, H., N. Beaumont, 2011: Green Tourism Futures: Climate Change Responses by Australian Government Tourism Agencies. In: CAUTHE 2011: National Conference: Tourism: Creating a Brilliant Blend, 8-11 February 2011, Adelaide, SA. pp 850-865. Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-1: Observed and projected changes in key climate variables, and (where assessed) the contribution of human activities to observed changes. For further relevant information see WGI Chapters 3, 6 (ocean changes, including acidification), 11, 12 (projections), 13 (sea level), and 14 (regional climate phenomena). (*) medium confidence, (**) high confidence, (***) very high confidence, (****) virtually certain Climate Observed change Direction of projected change Examples of projected magnitude of change Additional comments variable (relative to ~1990, unless otherwise stated) Mean air Aus: Increased by 0.09 +/- 0.03°C per decade Aus and NZ: Increase3-8 (****); greatest over Aus: 0.6-1.5°C (2030 A1B), 1.0-2.5°C (2070 Aus: A significant contribution to observed temperature since 19111 (***) inland Aus and least in coastal areas and NZ5-8 B1), 2.2-5.0°C (2070 A1FI)3 change attributed to anthropogenic climate NZ: Increased by 0.09 +/- 0.03°C per decade (***) NZ: 0.3-1.4°C (2040 A1B), 0.7-2.3°C (2090 change10 (**) with some regional variations since 19092 (***) B1), 1.6-5.1°C (2090 A1FI)5 attributed to atmospheric circulation variations11,12 CMIP5 RCP4.5, rel. to ~19959: N Aus: 0.3-1.6°C (2016-2035), 0.7-2.6°C NZ: Observed change partially attributed to (2046-2065) anthropogenic climate change13 (*) S Aus & NZ: 0.1-1.0°C (2016-2035), 0.6- 1.7°C (2046-2065) Sea surface Aus: Increased by about 0.12°C per decade for Aus and NZ: Increase3,7,8 (***) with greater Aus: 0.6-1.0°C (2070 B1) and 1.6-2.0°C temperature NW&NE Aus and by about 0.2°C per decade increase in the Tasman sea region (*) 3,7 (2070 A1FI) for southern coastal and 1.2- for SE Aus since 195014,15 (***) 1.5°C (2070 B1) and 2.2-2.5°C (2070 A1FI) NZ: Increased by about 0.07°C per decade elsewhere3 over 1909-20092 (***) NZ: Similar to projected changes in mean air temperature for coastal waters5 Air Aus and NZ: Significant trend since 1950: Aus and NZ: Hot days and nights more Aus: Hot days in Melbourne (>35°C max.) Aus: Observed trends partly attributable to temperature cool extremes have become rarer and hot frequent and cold days and cold nights less increase by 20-40% (2030 A1B), 30-90% anthropogenic climate change (**) as they are extremes extremes more frequent and intense16-19 frequent during the 21st century3,5,21-24 (**) (2070 B1) and 70-190% (2070 A1FI)3 consistent with mean warming and historical (**).The Australian heatwave of 2012/13 was NZ: Spring and autumn frost-free land to at simulations18,19,21,25, although other factors exceptional in heat, duration and spatial extent least triple by 2080s24; up to 60 more hot days may have contributed to high extremes during 20 . (>25°C max.) for northern areas by 20905 droughts26-28 Precipitation Aus: Late autumn/winter decreases in SW Aus Aus: Annual decline in SW Aus (**), Aus: For 2030 A1B, annual changes of-10% Aus: Observed decline in SW is related to since the 1970s and in SE Aus since the mid elsewhere on most of the southern (*) and NE to +5% (N Aus) and -10% to 0% (S Aus), for atmospheric circulation changes38-40 (***), 1990s, and annual increases in NW Aus since (low confidence) continental edges, with 2070 B1, -15% to +7.5% (N&E Aus) and - other factors41, and partly attributable to the 1950s29-31 (***) reductions strongest in the winter half 15% to 0% (S Aus), and for 2070 A1FI, -30% anthropogenic climate change40-43 (**). The NZ: Mean annual rainfall increased over year3,4,9,33-35 (**). Direction of annual change to +20% (N&E Aus) and -30% to +5% (S recent SE rainfall decline is also related to 1950-2004 in the south and west of the South elsewhere is uncertain3,35,36 (Figure 25.1) (**) Aus), with larger changes seasonally3 circulation changes31,44-46 (**), with some Island and west of the North Island, and NZ: In the South Island, annual increase in the NZ: For 2040 A1B, annual changes of -5% to evidence of an anthropogenic component47 decreased in the north-east of the South Island west and south and decrease in north-east. In +15% (S&W) and -15% to +10% (N&E) and NZ: Observed trends related to increased and east and north of the North Island32 (***) the North Island, increase in the west and for 2090 A1B, -10% to +25% (S&W) and - westerly winds32. Projected annual trends decrease in eastern and northern regions 5,34,37 20% to +15% (N&E) based on downscaled dominated by winter and spring trends related (Figure 25.1) (*) projections with larger changes seasonally5,37 to increased westerlies5 Precipitation Aus: Indices of annual daily extremes (e.g. Aus and NZ: Increase in most regions in the Aus: For 2090 A2, CMIP3 give increases in Aus and NZ: The sign of observed trends extremes 95th and 99th percentile rainfalls) show mixed intensity of rare daily rainfall extremes (i.e. the intensity of the 20 year daily extreme of mostly reflects trends in mean rainfall (e.g. or insignificant trends21,48, but significant current 20 year return period events) and in around +200% to -25% depending on region there is a decrease in mean and daily extremes increase is evident in recent decades for short duration (sub-daily) extremes (*) and an and model52 in SW Aus)21,32,49. Similarly, future increases shorter duration (sub-daily) events49,50 (**) increase in the intensity of 99 percentile daily NZ: Increases of daily extreme rainfalls of in intensity of extreme daily rainfall are more NZ: Extreme annual 1-day rainfall decrease in extremes (low confidence)5,8,21,51-56 around 8% per degree C are projected but with likely where mean rainfall is projected to north and east and increase in west since significant regional variations5,56 increase3,5 193032 (*) Subject to Final Copyedit 85 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Drought Aus: Defined using rainfall only, drought Aus and NZ: Drought frequency is projected Aus: Occurrence under 2070 A1B and A2 Aus: Regional warming may have led to an occurrence over the period 1900-2007 has not to increase in southern Australia 8,54,57,59,60 (*) ranges from a halving to 3 times more increase in hydrological drought (low changed significantly57 (**) and in many regions of New Zealand58,61 (*) frequent in N. Aus, and 0-5 times more confidence)62,63 NZ: Defined using a soil water balance model, frequent in southern Aus60 there has been no trend in drought occurrence NZ: Time spent in drought in eastern and since 197258 (*) northern New Zealand is projected to double or triple by 204061 Winds Aus: Significant decline in storminess over SE Aus: Increases in winds in 20-30°S band, with Aus: Magnitude of simulated mean changes Aus and NZ: Many of past and projected Aus since 188564 (*), but inconsistent trends in little change to decrease elsewhere, except for may exceed 10% under A1B for 2081-2100 changes in mean wind speed can be related to wind observations since 197565,66 winter increases over Tasmania. Decrease to relative to 1981-2000 69 changes in atmospheric circulation43,67,68 NZ: Mean westerly flow increased during the little change in extremes (99th percentile) over NZ: Mean westerly flow to increase by NZ: Extreme westerlies and southerlies have late 20th century (1978 1998), associated with most of Australia except Tasmania in winter69 around 20% in spring and around 70% in slightly increased while extreme easterlies the positive phase of the IPO67,68 (*) winter, and to decrease by around 20% in have decreased since 196013,71 NZ: Mean westerly winds and extreme winds summer and autumn, by 20905 (based on projected changes in circulation patterns) are projected to increase, especially in winter5,70 (*) Mean sea level Aus: From 1900-2011 the average rate of Aus and NZ: Regional sea level rise will very Aus: Off shore regional sea level rise may Aus and NZ: Satellite estimates of regional relative sea level rise (SLR) was 1.4+/-0.6 likely exceed the 1971-2000 historical rate, exceed 10% more than global SLR, see AR5 SLR for 1993-2009 are significantly higher mm/yr72 (***) consistent with global mean trends74. Mean WGI Chap13, Figure 13.2174 than those for 1920-2000, partly reflecting NZ: The average rate of relative SLR was sea level will continue to rise for at least NZ: Off shore regional sea level rise may be climatic variability72,73,76,77 1.7+/-0.1 mm/yr over 1900-200973 (***) several more centuries74 (***) up to 10% more than global SLR75 NZ: Allowing for glacial isostatic adjustment, absolute observed SLR is around 2.0mm/yr73,78 Extreme sea Aus and NZ: Extreme sea levels have risen at Aus and NZ: Projected mean SLR will lead to Aus: An increase of mean sea level by 0.1m level a similar rate to global SLR79 large increases in the frequency of extreme sea increases the frequency of an extreme sea level events (***), with other changes in storm level event by a factor of between 2 and 10 surges playing a lesser role80-83 over southeastern Australia depending on location 80-82, Fire weather Aus: Increased since 1973(**) with 24 out of Aus: Fire weather is expected to increase in Aus: Increase in days with very high and Aus: For the example of Canberra, the 38 sites showing increases in the 90th most of southern Australia due to hotter and extreme fire danger index by 2-30% (2020), 5- projected changes represent the current 17 percentile of the McArthur Forest Fire Danger drier conditions (**), based on explicit model 100% (2050) (using B1 and A2 and two days per year increasing to 18-23 days in 2020 index84 studies carried out for SE Australia 85-88, and climate models, and 1973-2007 base)85 and 20-33 days in 205085 change little or decrease in NE88 (*) NZ: Increase in days with very high and NZ: Fire danger index is projected to increase extreme fire danger index from around 0 to in many areas89 (*) 400% (2040) and 0 to 700% (2090) (using A1B,16 CMIP3 GCMs )89 Tropical Aus: No regional change in the number of Aus: Tropical cyclones are projected to Aus: Modelling study shows a 50% reduction Aus: Regional research on convective storms cyclones and tropical cyclones (TCs) or in the proportion of increase in intensity and stay similar or in TC occurrence for 2051-2090 relative to is limited but studies have shown a projected other severe intense TCs over 1981-200790 (*), but decrease in numbers9,94, and occur further 1971-2000, increases in intensity of the decrease in the frequency of cool-season storms frequency of severe landfalling TCs in NE south94 (low confidence) modelled storms, and occur around 100km tornadoes95, and hail3 in southern Australia, Aus has declined significantly since the late NZ: Projected increase in the average intensity further south94 and increases in the frequency and intensity of 19th Century91 and east-west distribution of cyclones in the south during winter, but a NZ: Occurrence of conditions conducive to hail in the Sydney region3,96 changed since 1980.92 decrease elsewhere70 (*) convective storm development is projected to There has been no trend in environments increase by 3 6% by 2070-2100 (A2), relative suitable for severe thunderstorms93 to 1970-2000, with the largest increases over the South Island70 Subject to Final Copyedit 86 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Snow and ice Aus: Late season significant snow depth Aus: Both snow depth and area are projected Aus: Area with at least 30 days cover annually NZ: Atmospheric circulation variations can decline at three out of four Snowy mountain to decline97 (***) projected to decline 14-54% (2020) and 30- enhance or outweigh multi-decadal trends in sites over 1957-200297 (**) NZ: Snowline elevations are projected to rise, 93% (2050)97 ice volume over time scales of up to two NZ: Ice volume declined by 36-61% from the and winter snow volume and days with low NZ: By 2090, peak snow accumulation is decades104,105 mid-late 1800s to the late 1900s98-100, with elevation snow cover are projected to projected to decline by 32-79% at 1000m and glacier volume reducing by 15% between decrease5,102,103 (***) by 6-51% at 2000 m103 1976 and 2008101 (**) References: 1 Fawcett et al. (2012); 2 Mullan et al. (2010); 3 CSIRO and BoM (2007); 4 Moise and Hudson (2008); 5 MfE (2008b); 6 AR5-WGI-Atlas-AI68-69; 7 AR5-WGI-Ch11; 8 AR5-WGI-Ch12; 9 AR5-WGI-Ch14; 10 Karoly and Braganza (2005); 11 Hendon et al. (2007); 12 Nicholls et al. (2010); 13 Dean and Stott (2009); 14 Lough (2008); 15 Lough and Hobday (2011); 16 Chambers and Griffiths (2008); 17 Gallant and Karoly (2010); 18 Nicholls and Collins (2006); 19 Trewin and Vermont (2010); 20 BOM (2013); 21 Alexander and Arblaster (2009); 22 Tryhorn and Risbey (2006); 23 Griffiths et al. (2005); 24 Tait (2008); 25 Alexander et al. (2007); 26 Deo et al. (2009); 27 McAlpine et al. (2007); 28 Cruz et al. (2010); 29 Hope et al. (2010); 30 Jones et al. (2009); 31 Gallant et al. (2012); 32 Griffiths (2007); 33 Timbal and Jones (2008); 34 AR5-WGI-Atlas-AI70-71; 35 Irving et al. (2012); 36 Watterson (2012); 37 Reisinger et al. (2010); 38 Bates et al. (2008); 39 Frederiksen and Frederiksen (2007); 40 Hope et al. (2006); 41 Timbal et al. (2006); 42 Cai and Cowan (2006); 43 Frederiksen et al. (2011); 44 Cai et al. (2011); 45 Nicholls (2010); 46 Smith and Timbal (2010); 47 Timbal et al. (2010a); 48 Gallant et al. (2007); 49 Westra and Sisson (2011); 50 Jakob et al. (2011); 51 Abbs and Rafter (2009); 52 Rafter and Abbs (2009); 53 Kharin et al. (2013); 54 IPCC-SREX-Chapter-3; 55 Westra et al. (2013); 56 Carey- Smith et al. (2010); 57 Hennessy et al. (2008a); 58 Mullan et al. (2005); 59 Kirono and Kent (2010); 60 Kirono et al. (2011); 61 Clark et al. (2011); 62 Cai and Cowan (2008); 63 Nicholls (2006); 64 Alexander et al. (2011); 65 McVicar et al. (2008); 66 Troccoli et al. (2012); 67 Parker et al. (2007); 68 Mullan et al. (2001); 69 McInnes et al. (2011a); 70 Mullan et al. (2011); 71 Salinger et al. (2005); 72 Burgette et al. (2013); 73 Hannah and Bell (2012); 74 AR5-WGI-Ch13; 75 Ackerley et al. (2013); 76 CSIRO and BoM (2012); 77 Meyssignac and Cazenave (2012); 78 Hannah (2004); 79 Menendez and Woodworth (2010); 80 McInnes et al. (2009); 81 McInnes et al. (2011b); 82 McInnes et al. (2012); 83 Harper et al. (2009); 84 Clarke et al. (2012); 85 Lucas et al. (2007); 86 Hasson et al. (2009); 87 Cai et al. (2009a); 88 Clarke et al. (2011); 89 Pearce et al. (2011); 90 Kuleshov et al. (2010); 91 Callaghan and Power (2011); 92 Hassim and Walsh (2008); 93 Allen and Karoly (2013); 94 Abbs (2012); 95 Timbal et al. (2010b); 96 Leslie et al. (2008); 97 Hennessy et al. (2008b); 98 Hoelzle et al. (2007); 99 Ruddell (1995); 100 Chinn (2001); 101 Chinn et al. (2012); 102 Fitzharris (2004); 103 Hendrikx et al. (2012); 104 Purdie et al. (2011); 105 Willsman et al. (2010). Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-2: Constraints and enabling factors for institutional adaptation processes in Australasia.* Constraint Enabling factors Uncertainty of projections Improved guidance and tools to manage uncertainty and support adaptive management 1-8 Increased focus on lead and consequence time of decisions and link with current climate variability and related risks 9-13 Increased communication between practitioners and scientists to identify and provide decision-relevant data and context 2,3,11,13-17 Availability and cost of data Central provision of relevant core climate and non-climate data, including regional and models scenarios of projected changes 4,5,7,9,18,19 National first-pass risk assessments 4,5,7,8,18,20-24 Limited financial and human Support for pilot projects 4,8,15,18,24,25 capability and capacity; time Building capacity through institutional commitment and learning 3,5,11,17,23,26-28 lag in developing expertise Central databases on guidance, tools, methodologies, case studies 4,5,7,18,24 Regional partnerships and collaborations, knowledge networks 3,4,8,13,15,17,26,28-30 Unclear problem definition Explicit but iterative framing and scoping of adaptation challenge, to reflect alternative and goals; unclear standards entry points for stakeholders while meeting expectations of project sponsors to ensure for risk assessment long-term support 3,11,17,31-34 methodologies and decision Tailoring decision-making frameworks to specific problems 1,2,6,17,35,36 support tools; limited Criteria and tools to monitor and evaluate adaptation success 7,18,37-39 monitoring and evaluation Unclear or contradictory Clear and coordinated legislative frameworks 5,8,9,15,24,40-45 legislative frameworks and Defined responsibilities for public and private actors, including liabilities from acting and responsibilities, unclear failure to act 8,9,11,24,41,44,46 liabilities Legally binding guidance on the incorporation of climate change in planning mechanisms 5,7,8,15,38,40 Static planning mechanisms Whole-of-council approach to climate adaptation to break up institutional and professional and practice; competing silos 15,33,47 mandates and fragmentation Long-term policy commitments and implementation support 5,18,26,33,48 of policies; disciplinary voids Increased policy coherence across sectors, regulations and levels of government 9,26,28,40,42,43,47 or single approaches Enabling risk-based flexible land-use decisions 4,5,9,49 Strengthening multi-disciplinarity across professional fields 14,29,48 Lack of political leadership; Legally binding guidance and clarification of liabilities and duty of care to reduce short election cycles; limited dependence on individual leadership 5,7-9,15,24,38,40,46,49 community support, Consistent but audience-specific communication of current and potential future participation and awareness vulnerability and implications for community values 4,5,7,26,42,43,50 for adaptation Comprehensible communication of and access to response options, and their consistency with wider development plans 7,26,28,33,39,42,43 Clearly identified entry points for public participation 17,34,38,39,42,48,51-53 * Note: The relevance of each constraint varies among organisations, sectors and location. Some enabling factors are only beginning to be implemented or have only been suggested in the literature, hence their effectiveness cannot yet be evaluated. Entries for enabling factors exclude generic mechanisms, such as insurance (see Box 25-7), emergency management and early warning systems, and funding for pilot studies, capital infrastructure upgrades or retreat schemes. References: 1 Randall et al. (2012); 2 Verdon-Kidd et al. (2012); 3 Webb et al. (2013); 4 Mukheibir et al. (2013); 5 Lawrence et al. (2013b); 6 Nelson et al. (2008); 7 Britton (2010); 8 Gurran et al. (2008); 9 Productivity Commission (2012); 10 Stafford-Smith et al. (2011); 11 Johnston et al. (2013); 12 Park et al. (2012); 13 Power et al. (2005); 14 Reisinger et al. (2011); 15 Smith et al. (2008); 16 Stafford-Smith (2013); 17 Yuen et al. (2012); 18 Webb and Beh (2013); 19 Roiko et al. (2012); 20 DCCEE (2011); 21 DCC (2009); 22 Baynes et al. (2012); 23 Smith et al. (2010); 24 SCCCWEA (2009); 25 DSEWPC (2011); 26 Low Choy et al. (2012); 27 Gardner et al. (2010); 28 Fidelman et al. (2013); 29 Mustelin et al. (2013); 30 Serrao-Neumann et al. (2013); 31 Fünfgeld et al. (2012); 32 Kuruppu et al. (2013); 33 Britton et al. (2011); 34 Alexander et al. (2012); 35 Maru et al. (2011); 36 Preston et al. (2008); 37 Norman et al. (2012); 38 Rouse and Norton (2010); 39 Preston et al. (2011); 40 Rive and Weeks (2011); 41 Abel et al. (2011); 42 Norman (2009); 43 Gurran et al. (2006); 44 McDonald (2013); 45 Minister of Conservation (2010); 46 McDonald (2010); 47 Measham et al. (2011); 48 Rouse and Blackett (2011); 49 McDonald (2011); 50 Hine et al. (2013); 51 Burton and Mustelin (2013); 52 Hobson and Niemeyer (2011); 53 Gardner et al. (2009a). Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-3: Examples of detected changes in species, natural and managed ecosystems, consistent with a climate change1 signal, published since the AR4. Confidence in detection of change is based on the length of study, and the type, amount and quality of data in relation to the natural variability in the particular species or system. Confidence in the role of climate being a major driver of the change is based on the extent to which the detected change is consistent with that expected under climate change, and to which other confounding or interacting non-climate factors have been considered and been found insufficient to explain the observed change. Type of change and nature Examples Time scale of Confidence in the Potential climate change Confidence in the role of evidence observations detection of driver(s) 1 of climate vs other biological change drivers Morphology Declining body size of southeast Australian ~100 years medium Warming air temperatures medium passerine birds, equivalent to ~7o latitudinal shift trend significant for 4 ~1.0oC over same period Nutritional cause Limited evidence (Gardner et al., 2009) out of 8 species, two discounted (1 study) other species show same trend but not statistically significant Geographic distribution Southerly range extension of the barrens-forming ~30-50 years high Increased sea surface high sea urchin Centrostephanus rodgersii from the New (first recorded in temperature (SST), Ocean High agreement, robust South Wales coast to Tasmania; flow on impacts to Tasmania late warming in SE Australia, evidence for many marine marine communities including lobster fishery; shift 1970s) increased southerly penetration species & mobile terrestrial of 160 km per decade over 30 years (Ling, 2008; of the East Australian Current species Ling et al., 2008; Ling et al., 2009; Banks et al., (EAC), 350 km over 60 years 2010) Forty-five fish species, representing 27 families distributions high Increased SST SE Australia, medium (about 30% of the inshore fish families occurring in from late 1880s, increased southerly penetration Changed fishing the region), exhibited major distributional shifts in 1980s and of EAC practices have potentially Tasmania (Last et al., 2011) present (1995- contributed to trends now) Southward range shift of intertidal species (average ~50 years medium Increased SST in SE Australia medium minimum distance 116 km) off west coast of Sites resampled (average 0.22oC per decade), Tasmania; 55% species recorded at more southerly 2007-2008, increased southerly penetration sites, only 3% species expanded to more northerly compared with of the EAC, 350 km over 60 sites (Pitt et al., 2010) 1950s years Subject to Final Copyedit 89 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Life cycles Significant advance in mean emergence date of 1.5 65 years high Increase in local air high days per decade (1941-2005) in the Common temperatures of 0.16oC per Advance consistent with Robust evidence, medium Brown butterfly Heteronympha merope in Australia decade (1945-2007) physiologically based agreement; increasing (Kearney et al., 2010) model of temperature documentation of advances influence on in phenology in some development species (mainly migration Advances in spring phenology of migratory birds, Multiple time high Local climate trends high and reproduction in birds, and both advances and delays in phenology in other periods from (increasing air temperature, No other potential emergence in butterflies, seasons at multiple Australian sites: meta-analysis 1960s, all decreased raindays) were more confounding facotrs flowering in plants) but also of 52 species and 145 datasets (Chambers et al., included 1990s important than broad-scale identified significant trends towards 2013b) and 2000s drivers such as the Southern later life cycle events in Oscillation Index. Strongest some taxa (see meta-analysis associations were with for Southern Hemisphere decreased raindays. phenology (Chambers et al., Earlier wine-grape ripening at 9 of 10 sites in Multiple time high Increased length of growing medium 2013a) Australia (Webb et al., 2012) periods up to 64 season, increased average Changed husbandry years (average 41 temperature and reduced soil techniques, resulting in years) moisture lower crop yields, may have contributed to trend Timing of migration of glass eels, Anguilla spp. 30 years medium Warming water temperatures in low advanced by several weeks in Waikato River, North (2004-2005 spawning grounds Changes in discharge island, New Zealand (Jellyman et al., 2009) compared to discounted as 1970s) contributing factor Marine productivity Otolith ( ear stone ) analyses in long-lived Pacific Birth years high Increasing growth rates in medium fish indicates significantly increased growth rates ranged 1861- species in top 250m associated Changed fishing pressure Limited evidence, medium for shallow-water species (<250 m) (3 of 3 1993 (fish 2-128 with warming SST, declining may have contributed to agreement species), reduced growth rates of deep-water years old) growth rates in species >1000m trend (>1000 m) species (3 of 3 species); no change associated with long-term observed in the 2 intermediate-depth species cooling (as indicated by Mg/Ca (Thresher et al., 2007) ratios and delta18O in deep water corals) ~50% decline in growth rate and biomass of spring 60 year dataset; high Increased SST and extension medium phytoplankton bloom in western Tasman Sea decline recorded EAC associated with reduced (Thompson et al., 2009) over period nutrient availability 1997-2007 Subject to Final Copyedit 90 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Vegetation change Expansion of monsoon rainforest at expense of ~40 years medium Increases in rainfall and medium eucalypt savanna and grassland in Northern atmospheric CO2 Changes in fire regimes Limited agreement & Territory, Australia (Banfai and Bowman, 2007; and land management evidence; interacting impacts Bowman et al., 2010) practices may have of changed land practices, contributed to trend altered fire regimes, Net increase in mire wetland extent (10.2%) and Weather data medium Decline in evapo-transpiration low increasing atmospheric CO2 corresponding contraction of adjacent eucalypt covers >40 years Resource exploitation, concentration and climate woodland in seven sub-catchments in south east (depending on fire history and autogenic trends difficult to Australia (Keith et al., 2010) parameter); mire development disentangle vegetation discounted mapping from 1961-1998 Freshwater communities Decline in families of macroinvertebrates that 13 years medium Increasing water temperatures low favour cooler, faster-flowing habitats in New South (1994-2007) and declining flows Variation in sampling, Limited evidence Wales streams and increase in families favouring changes in water quality, (1 study) warmer and more lentic conditions (Chessman, impacts of impoundment 2009) and water extraction may have contributed to trends Disease Emergence and increased incidence of coral 1998 onwards medium Increasing SST high diseases including white syndrome (since 1998), Limited evidence, robust and black band disease (since 1993-4) (Bruno et al., agreement 2007; Sato et al., 2009; Dalton et al., 2010) Coral reefs Multiple mass bleaching events since 1979 (see 1979 onwards high Increasing SST high 25.6.2, 30.5) Robust evidence & high Calcification of Porites on GBR declined 21% 1971-2003; high Increasing SST high agreement (1971-2003, 4 reefs; Cooper et al., 2008); about 1990-2005 Changes in water quality 11% (1990-2005, 69 reefs; De'ath et al., 2009) discounted Subject to Final Copyedit 91 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-4: Examples of potential consequences of climate change for invasive and pathogenic species relevant to Australia and New Zealand, with consequence categories based on Hellman et al. (2008). Consequence Projected change Organism/Ecosystem affected altered mechanisms Increased risk of introduction of Asiatic Citrus Psyllid, (Diaphorina citri), vector of the disease huanglongbing 1 Australian citrus industry and native citrus and other rutaceaous of transport and species and endemic psyllid fauna introduction altered distribution Nassella neesiana (Chilean needle grass): increased droughts favour establishment 2 Managed pasture in New Zealand of existing invasive Warming and drying may encourage the spread of existing invasives such as Pheidole megacephala in New Human health and potentially agricultural and natural & pathogenic species Zealand and provide suitable conditions for other exotic ant species if they invade 3 ecosystems Reduced climatic suitability for exotic invasive grasses in Australia (11 species including Nassella sp.) 4 Range of the invasive weed Lantana camara (lantana) projected to extend from Northern Australia to Victoria, Australian rangeland South Australia and Tasmania 5 Projected increases in the range of three recently naturalised sub-tropical plants (Archontophoenix Multiple cunninghamiana, Psidium guajava, Schefflera actinophylla) 6 Native ecosystems in New Zealand altered climatic Queensland fruit fly (Bactrocera tryoni) moving southwards 7 Australian horticulture constraints on Significant association between amphibian declines in upland rainforests of north Queensland and three Native frogs invasive & consecutive years of warm weather suggests future warming could increase the vulnerability of frogs to pathogenic species chytridiomycosis caused by the chytrid fungus Batrachochytrium dendrobatadis 8 altered impact of Fusarium pseudograminearum causing crown rot increases under elevated CO2 9 Australian wheat existing invasive & Increased abundance of the root-feeding nematode Longidorus elongatus under elevated CO2 10 New Zealand pasture pathogenic species Increased severity of Swiss needle cast disease caused by Phaeocryptopus gaeumannii 11 Douglas fir plantations in New Zealand, impact more severe in North Island altered effectiveness Light brown apple moth, Epiphyas postvittana (Walker) (Lepidoptera:Tortricidae) reduction in natural enemies Australian horticulture of management due to asynchrony and loss of host species 12 strategies Projected changes in the efficacy of five biological control systems demonstrating a range of potential disruption Pastoral and horticultural systems in New Zealand mechanisms 13 References: 1 Finlay et al. (2009); 2 Bourdôt et al. (2012); 3 Harris and Barker (2007); 4 Gallagher et al. (2012a); 5 Taylor et al. (2012b); 6 Sheppard (2012); 7 Sutherst et al. (2000); 8 Laurance (2008); 9 Melloy et al. (2010); 10 Yeates and Newton (2009); 11 Watt et al. (2011b); 12 Thomson et al. (2010); 13 Gerard et al. (2012). Subject to Final Copyedit 92 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-5: Examples of co-beneficial climate change adaptation options for urban areas and barriers to their adoption. Options in italics are already widely implemented in Australia and New Zealand urban areas. Climate impact Adaptation options Co-benefits Barriers to adoption Hot days and heatwaves 1-8 Greening cities/roofs; more green spaces; Energy efficiency; reduced risk of Lack of standards; high installation costs; well-designed energy efficient buildings; blackouts; fewer health impacts; limited understanding of benefits; high occupant behavioral change; standards for resilient infrastructure and assets; individual discount rate; split of private costs new and retrofitting of existing resilient community and public benefits infrastructure and assets; new methods and material for transport infrastructure to withstand higher extreme temperature Decreased water supply and Supply augmentation (water recycling, Water self-sufficiency for current and Potential health impacts of recycled water; drought rainwater harvesting, increased storage, future demand/population; less lower than expected uptake of demand desalinisation); demand management; pipe/storage leakage; reduced options and relaxation after crises; trade-offs [See Box 25-2 for more] infrastructure upgrades; integrated urban environmental impacts from abstraction between supply and demand management; sensitive design cost and environmental impacts of some augmentation options River and local flooding, New standards and improvements to Reduced damages to homes and High implementation cost especially if coastal erosion and building, water infrastructure (e.g. infrastructure and loss of life; decreased retrospective on existing stock; rezoning/ inundation drainage and sewerage) and transport insurance premiums; habitat protection relocation can affect property prices and are infrastructure; upgrades of protection highly contested [See Boxes 25-1 and 25-8 systems; retaining floodplains/floodways; for more] restoring wetlands; buffers from hazard- prone areas; raising minimum floor levels; rezoning/ relocation Severe storms and tropical New building design to withstand higher Reduced damages to homes and High implementation cost; rezoning/ cyclones 9-12 wind pressures; rezoning/relocation infrastructure and loss of life; decreased relocation can affect property prices and are insurance premiums highly contested Corrosion from increased Improved standards for construction Reduced rates of carbonation-induced Effectiveness of coatings varies with age and atmospheric CO2 levels 13,14 using concrete; application of coatings for corrosion of concrete condition of concrete existing building stock References: 1 BRANZ (2007); 2 Coutts et al. (2010); 3 Moon and Han (2011); 4 Stephenson et al. (2010); 5 Williams et al. (2010); 6 CSIRO et al. (2007); 7 Taylor and Philp (2010); 8 QUT (2010); 9 Mason and Haynes (2010); 10 Wang et al. (2010b); 11 Stewart and Wang (2011); 12 Mason et al. (2013); 13 Stewart et al. (2012); 14 Wang et al. (2012). Subject to Final Copyedit 93 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-6: Examples of interactions between impacts and adaptation measures in different sectors. In each case, impacts or responses in one sector have the potential to conflict (cause negative impacts) or be synergistic (have co-benefits) with impacts or responses in another sector, or with another type of response in the same sector. Primary goal Sector(s) Examples of interactions between impacts and adaptation responses affected Reduction of bushfire risk Biodiversity, Potential for greater conflict between conservation managers and other park users in Kosciuszko National Park if increasing fire in natural landscapes tourism incidence causes park closures, either to reduce risk, or to rehabilitate vegetation after fires (Wyborn, 2009), e.g. Objectives of the Wildfire Management Overlay (WMO) in Victoria conflicts with vegetation conservation (Hughes and Mercer, 2009). Reduction of risk to energy Biodiversity, Underground cabling would reduce both the susceptibility of transmission networks to fire and ignition sources for wild fires, thus transmission from energy reducing risks to ecosystems and settlements; constraints include significant investment cost, diverse ownership of assets and lack bushfires of an overarching national strategy (ATSE, 2008; Parsons Brinkerhoff, 2009; Linnenluecke et al., 2011). Protection of coastal Biodiversity, Seawalls may provide habitat but these communities have different diversity and structure to those developing on natural substrates infrastructure tourism (Jackson et al., 2008); groynes potentially alter beach fauna diversity and community structure (Walker et al., 2008); continuing hard protection against sea level rise results in long-term loss of coastal amenities (Gorddard et al., 2012). Avoidance of risks from sea Indigenous Relocation can avoid increasing local pressures on communities from sea level rise but raises complex cultural, land rights, legal level rise via relocation communities and economic issues, e.g. potential relocation of Torres Strait islander communities (Green et al., 2010b; McNamara et al., 2011). Allocating scarce water Rural areas, Market based instruments such as water trading help allocation of scarce water resources to the highest value uses. The negative resources via market agriculture, implications of this include potential loss of access to lower value users, which in some areas includes agriculture and drinking instruments mining water supplies, with potentially significant social, environmental and wider economic consequences (Kiem and Austin, 2012). Increased water security Biodiversity, Water storage can buffer urban settlements and agricultural systems against high variability in river flows, but altered flow regimes via augmentation of supply water demand can have significant negative impacts on freshwater ecosystems (Bond et al., 2008; Pittock et al., 2008; Kingsford, 2011). for urban and agricultural management Discharge from desalination plants (e.g. in Perth and Sydney) can lead to substantial local increases in salinity and temperature, systems and the accumulation of metals, hydrocarbons and toxic anti-fouling compounds in receiving waters (Roberts et al., 2010); increasing supply can reduce the effectiveness of demand-side measures (Barnett and O'Neill, 2010; Taptiklis, 2011; Box 25-2). Subject to Final Copyedit 94 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-7: Examples of interactions between adaptation and mitigation measures (green rows denote synergies where multiple benefits may be realized, orange rows denote potential tradeoffs and conflicts; grey row gives an example of complex, mixed interactions). The primary goal may be adaptation or mitigation. Primary goal Sector(s) Examples of interactions between adaptation and mitigation responses affected Adaptation to Biodiversity, Snowmaking in the Australian Alps would require large additional energy and water resources by 2020 of 2500-3300 ML of water per decreasing snowfall energy use, month, more than half the average monthly water consumption by Canberra in 2004-05. Increased snowmaking negatively affects water use vegetation, soils and hydrology of subalpine-alpine areas (Pickering and Buckley, 2010; Morrison and Pickering, 2011; ABS, 2012a). Air conditioning for Health, Rising temperatures degrade building energy efficiency (Wang et al., 2010a) and increase energy demand and associated CO2 emissions heat stress energy use if summer cooling needs are met by increased air conditioning (Stroombergen et al., 2006; Thatcher, 2007; Wang et al., 2010a). Renewable wind Biodiversity Wind-farms can have localised negative effects on bats and birds. However, risk assessment of the potential negative impacts of wind energy production turbines on threatened bird species in Australia indicated low to negligible impacts on all species modelled (Smales, 2006). Urban densification Biodiversity, Higher urban density to reduce energy consumption from transport and infrastructure can result in loss of permeable surfaces and tree water, health cover, intensify flood risks, and exacerbate discomfort and health impacts of hotter summers (Hamin and Gurran, 2009). Water supply from Energy Meeting increasing urban water demand via desalination plants increases energy demand and CO2 emissions if this demand is met by desalination demand increased fossil fuel energy generation (Barnett and O'Neill, 2010; Stamatov and Stamatov, 2010). Secure food Nitrous oxide Net greenhouse gas emissions intensity from dairy systems in southern Australia have been estimated to increase in future in several production in a and methane locations due to a changing climate and management responses (Cullen and Eckard, 2011; Eckard and Cullen, 2011). A shift towards warming climate emissions perennial C4 grasses would increase methane emissions from grazing ruminants due to lower feed quality, but studies in south-west Australia suggest this could be more than offset by increased soil carbon storage (Thomas et al., 2012; Bradshaw et al., 2013). Housing design to Energy use, Reducing peak energy demand through building design and demand management reduces vulnerability of electricity networks and reduce peak energy infrastructure, transmission losses during heat waves (Parsons Brinkerhoff, 2009; Nguyen et al., 2010), reduces heat stress during summer and provides demand health health benefits during winter (Strengers, 2008; Howden-Chapman, 2010; Strengers and Maller, 2011; Ren et al., 2012). Energy from Biodiversity, New crops such as oil mallees or other eucalypts may provide multiple benefits, especially in marginal areas, displacing fossil fuels or second-generation rural areas, sequestering carbon, generating income for landholders (essential oils, charcoal, bio-char, biofuels), and providing ecosystem services biofuels agriculture including reducing erosion (Cocklin and Dibden, 2009; Giltrap et al., 2009; McHenry, 2009). Reduced emissions Biodiversity, Improved management of savanna fires to reduce the extent of high intensity late season fires could substantially reduce emissions as from fires livelihoods well as having significant benefits for biodiversity and indigenous employment (Russell-Smith et al., 2009; Bradshaw et al., 2013). Reduce methane Biodiversity, Feral camels in Australia are projected to double from 1 to 2 million by 2020. Controlling their numbers to reduce methane emissions emissions from agriculture could have significant biodiversity benefits (NRMMC, 2010; Bradshaw et al., 2013). Economic benefits of reduced grazing competition, feral camels infrastructure damage and greenhouse gases could outweigh costs of camel reductions (Drucker et al., 2010). Subject to Final Copyedit 95 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 25-8: Key regional risks from climate change and the potential for reducing risk through mitigation and adaptation. Key risks are identified based on assessment of the literature and expert judgments by chapter authors, with evaluation of evidence and agreement in the supporting chapter sections. Each key risk is characterized on a scale from very low to very high and presented in three timeframes: the present, near-term (2030-2040), and long-term (2080-2100). For the near-term era of committed climate change (here, for 2030-2040), projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. For the longer-term era of climate options (here, for 2080- 2100), risk levels are presented for global mean temperature increase of 2°C and 4°C above preindustrial levels. For each timeframe, risk levels are estimated for a continuation of current adaptation and for a hypothetical highly adapted state. Relevant climate variables are indicated by icons. For a given key risk, change in risk level through time and across magnitudes of climate change is illustrated, but because the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks, sectors, or regions. Subject to Final Copyedit 96 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 25-1: Observed and projected changes in annual average temperature and precipitation. (Top panel, left) Observed temperature trends from 1901-2012 determined by linear regression [WGI AR5 Figures SPM.1 and 2.21]. (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression [WGI AR5 Figure SPM.2]. For observed temperature and precipitation, trends have been calculated where sufficient data permits a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level. Diagonal lines indicate areas where change is not significant. (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and >90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in >66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5 [Boxes 21-3 and CC-RC]. Subject to Final Copyedit 97 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 25-2: Observed and simulated variations in past and projected future annual average near-surface air temperature over land areas of Australia (left) and New Zealand (right). Black lines show various estimates from observational measurements. Shading denotes the 5-95 percentile range of climate model simulations driven with historical changes in anthropogenic and natural drivers (63 simulations), historical changes in natural drivers only (34), the RCP2.6 emissions scenario (63), and the RCP8.5 (63). Data are anomalies from the 1986-2005 average of the individual observational data (for the observational time series) or of the corresponding historical all- forcing simulations. Further details are given in Box 21-3. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 98 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 25-3: Adaptation as an iterative risk management process. Individual adaptation decisions comprise well known aspects of risk assessment and management (top left panel). Each decision occurs within and exerts its own sphere of influence, determined by the lead- and consequence time of the decision, and the broader regulatory and societal influences on the decision (top right panel). A sequence of adaptation decisions creates an adaptation pathway (bottom panel). There is no single correct adaptation pathway, although some decisions, and sequences of decisions, are more likely to result in long-term maladaptive outcomes than others, but the judgment of outcomes depends strongly on societal values, expectations and goals. Subject to Final Copyedit 99 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 25-4: Estimated changes in mean annual runoff for 1°C global average warming above current levels. Maps show changes in annual runoff (percentage change; top row) and runoff depth (millimetres; bottom row), for dry, median and wet (10th to 90th percentile) range of estimates, based on hydrological modelling using 15 CMIP3 climate projections (Chiew et al., 2009; CSIRO, 2009; Petheram et al., 2012; Post et al., 2012). Projections for 2°C global average warming are about twice that shown in the maps (Post et al., 2011). (Figure adapted from Chiew and Prosser, 2011; Teng et al., 2012). Subject to Final Copyedit 100 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 25 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 25-5: Projected changes in exposure to heat under a high emissions scenario (A1FI). Maps show the average number of days with peak temperatures >40°C, for ~1990 (based on available meteorological station data for the period 1975-2004), ~2050 and ~2100. Bar charts show the change in population heat exposure, expressed as person- days exposed to peak temperatures >40°C, aggregated by State/Territory and including projected population growth for a default scenario. Future temperatures are based on simulations by the GFDL-CM2 global climate model (Meehl et al., 2007), re-scaled to the A1FI scenario; simulations based on other climate models could give higher or lower results. Data from Baynes et al. (2012). Subject to Final Copyedit 101 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 26. North America Coordinating Lead Authors Patricia Romero-Lankao (Mexico), Joel B. Smith (USA) Lead Authors Debra Davidson (Canada), Noah Diffenbaugh (USA), Patrick Kinney (USA), Paul Kirshen (USA), Paul Kovacs (Canada), Lourdes Villers Ruiz (Mexico) Contributing Authors William Anderegg (USA), Jessie Carr (USA), Anthony Cheng (USA), Thea Dickinson (Canada), Ellen Douglas (USA), Rob de Loë (Canada), Hallie Eakin (USA), Daniel M. Gnatz (USA), Mary Hayden (USA), Maria Eugenia Ibarraran Viniegra (Mexico), Elena Jiménez Cisneros (Mexico), Michael D. Meyer (USA), Amrutasri Nori-Sarma (India), Landy Sánchez Pena (Mexico), Catherine Ngo (USA), Greg Oulahen (Canada), Diana Pape (USA), Ana Pena del Valle (Mexico), Roger Pulwarty (USA), Ashlinn Quinn (USA), Daniel Runfola (USA), Fabiola S. Sosa- Rodrigquez (Mexico), Bradley H. Udall (USA), Fiona Warren (Canada), Kate Weinberger (USA), Tom Wilbanks (USA) Review Editors Ana Rosa Moreno (Mexico), Linda Mortsch (Canada) Volunteer Chapter Scientist William Anderegg (USA) Contents Executive Summary 26.1. Introduction 26.2. Key Trends Influencing Risk, Vulnerability, and Capacities for Adaptation 26.2.1. Demographic and Socioeconomic Trends 26.2.1.1. Current Trends 26.2.1.2. Future Trends 26.2.2. Physical Climate Trends 26.2.2.1. Current Trends 26.2.2.2. Climate Change Projections 26.3. Water Resources and Management 26.3.1. Observed Impacts of Climate Change on Water Resources 26.3.2. Projected Climate Change Impacts and Risks 26.3.2.1. Water Supply 26.3.2.2. Water Quality 26.3.2.3. Flooding 26.3.2.4 Instream Uses 26.3.3. Adaptation 26.4. Ecosystems and Biodiversity 26.4.1. Overview 26.4.2. Tree Mortality and Forest Infestation 26.4.2.1. Observed Impacts 26.4.2.2. Projected Impacts and Risks 26.4.3. Coastal Ecosystems Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.4.3.1. Observed Climate Impacts and Vulnerabilities 26.4.3.2. Projected Impacts and Risks 26.4.4. Ecosystems Adaptation, and Mitigation 26.5. Agriculture and Food Security 26.5.1. Observed Climate Change Impacts 26.5.2. Projected Climate Change Risks 26.5.3. A Closer Look at Mexico 26.5.4. Adaptation 26.6. Human Health 26.6.1. Observed Impacts, Vulnerabilities and Trends 26.6.1.1. Storm-Related Impacts 26.6.1.2. Temperature Extremes 26.6.1.3. Air Quality 26.6.1.4. Pollen 26.6.1.5. Waterborne Diseases 26.6.1.6. Vectorborne Diseases 26.6.2. Projected Climate Change Impacts 26.6.3. Adaptation Responses 26.7. Key Economic Sectors and Services 26.7.1. Energy 26.7.1.1. Observed Impacts 26.7.1.2. Projected Impacts 26.7.1.3. Adaptation 26.7.2. Transportation 26.7.2.1. Observed Impacts 26.7.2.2. Projected Impacts 26.7.2.3. Adaptation 26.7.3. Mining 26.7.3.1. Observed Impacts 26.7.3.2. Projected Impacts 26.7.3.3. Adaptation 26.7.4. Manufacturing 26.7.4.1. Observed Impacts 26.7.4.2. Projected Impacts 26.7.4.3. Adaptation 26.7.5. Construction and Housing 26.7.5.1. Observed Impacts 26.7.5.2. Projected Impacts 26.7.5.3. Adaptation 26.7.6. Insurance 26.7.6.1. Observed Impacts 26.7.6.2. Projected Impacts 26.7.6.3. Adaptation 26.8. Urban and Rural Settlements 26.8.1. Observed Weather and Climate Impacts 26.8.2. Observed Factors and Processes Associated with Vulnerability 26.8.2.1. Urban Settlements 26.8.2.2. Rural Settlements 26.8.3. Projected Climate Risks on Urban and Rural Settlements 26.8.4. Adaptation Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.8.4.1. Evidence of Adaptation 26.8.4.2. Opportunities and Constraints 26.9. Federal and Subnational Level Adaptation 26.9.1. Federal Level Adaptation 26.9.2. Subnational Level Adaptation 26.9.3. Barriers to Adaptation 26.9.4. Maladaptation, Trade-Offs, and Co-Benefits 26.10. Key Risks, Uncertainties, Knowledge Gaps, and Research Needs 26.10.1. Key Multi-Sectoral Risks 26.10.2. Uncertainties, Knowledge Gaps, and Research Needs References Chapter Boxes 26-1. Adapting in a Transboundary Context: the Mexico-U.S. Border Region 26-2. Wildfires 26-3. Climate Responses in Three North American Cities Frequently Asked Questions 26.1: What impact is climate having on North America? 26.2: Can adaptation reduce the adverse impacts of climate in North America? Executive Summary Overview North America s climate has changed and some societally-relevant changes have been attributed to anthropogenic causes (very high confidence) [Figure 26-1]. Recent climate changes and individual extreme events demonstrate both impacts of climate-related stresses and vulnerabilities of exposed systems (very high confidence) [Figure 26-2]. Observed climate trends in North America include an increased occurrence of severe hot weather events over much of the US, decreases in frost days, and increases in heavy precipitation over much of North America (high confidence). [26.2.2.1] The attribution of observed changes to anthropogenic causes has been established for some climate and physical systems (e.g., earlier peak flow of snowmelt run-off and declines in the amount of water stored in spring snowpack in snow-dominated streams and areas of western United States and Canada (very high confidence) [Figure 26-1]. Evidence of anthropogenic climatic influence on ecosystems, agriculture, water resources, infrastructure, and urban and rural settlements is less clearly established, though, in many areas, these sectors exhibit substantial sensitivity to climate variability (high confidence) (26.3.1; 26.3.2; 26.4.2.1; 26.4.2.2; 26.4.3.1; Box 26-3; 26.5.1; 26.7.1.1; 26.7.2; 26.8.1; Figure 26-2). Many climate stresses that carry risk particularly related to severe heat, heavy precipitation and declining snowpack will increase in frequency and/or severity in North America in the next decades (very high confidence). Global warming of approximately 2°C (above the pre-industrial baseline) is very likely to lead to more frequent extreme heat events and daily precipitation extremes over most areas of North America, more frequent low snow years, and shifts towards earlier snowmelt runoff over much of the western US and Canada [26.2.2.2]. Together with climate hazards such as higher sea levels and associated storm surges, more intense droughts, and increased precipitation variability, these changes are projected to lead to increased stresses to water, agriculture, economic activities and urban and rural settlements (high confidence) [26.3.2.1-26.3.2.4; 26.5.2; 26.7.1.2; 26.8.3]. Global warming of approximately 4°C is very likely to cause larger changes in extreme heat events, daily-scale precipitation extremes and snow accumulation and runoff, as well as emergence of a locally-novel temperature regime throughout North America [26.2.2.2]. This higher level of global temperature change is likely to cause decreases in annual precipitation over much of the southern half of the continent and increases in annual Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 precipitation over much of the northern half of the continent [26.2.2.2]. The higher level of warming would present additional and substantial risks and adaptation challenges across a range of sectors (high confidence). [26.3.3, 26.5.2, 26.6.2, 26.7.2.2, 26.8.3] We highlight below key findings on impacts, vulnerabilities, projections, and adaptation responses relevant to specific North American sectors: ecosystems, water, agriculture, human health, urban and rural settlements, infrastructure and the economy. We then highlight challenges and opportunities for adaptation, and future risks and adaptive capacity for three key climate-related risks. Sector-Specific Climate Risks and Adaptation Opportunities North American ecosystems are under increasing stress from rising temperatures, CO2 concentrations, and sea-levels, and are particularly vulnerable to climate extremes (very high confidence). Climate stresses occur alongside other anthropogenic influences on ecosystems, including land-use changes, non-native species, and pollution, and in many cases will exacerbate these pressures (very high confidence). [26.4.1; 26.4.3]. Evidence since the Fourth Assessment Report highlights increased ecosystem vulnerability to multiple and interacting climate stresses in forest ecosystems, through wildfire activity, regional drought, high temperatures, and infestations (medium confidence) [26.4.2.1; Box 26-2]; and in coastal zones due to increasing temperatures, ocean acidification, coral reef bleaching, increased sediment load in run-off, sea level rise, storms, and storm surges (high confidence) [26.4.3.1]. In the near term, conservation and adaptation practices can buffer against climate stresses to some degree in these ecosystems, both through increasing system resilience, such as forest management to reduce vulnerability to infestation, and in reducing co-occurring non-climate stresses, such as careful oversight of fishing pressure (medium confidence) [26.4.4]. Water resources are already stressed in many parts of North America due to non-climate change anthropogenic forces, and are expected to become further stressed due to climate change (high confidence) [26.3, 26.3.1]. Decreases in snowpacks are already influencing seasonal streamflows (high confidence) [26.3.1]. While indicative of future conditions, recent floods, droughts, and changes in mean flow conditions cannot yet be attributed to climate change (medium to high confidence) [26.3.1, 26.3.2]. The 21st century is projected to witness decreases in water quality and increases in urban drainage flooding throughout most of North America under climate change as well as a decrease in instream uses such as hydropower in some regions (high confidence) [26.3.2.2, 26.3.2.3, 26.3.2.4]. Additionally, there will be decreases in water supplies for urban areas and irrigation in North America except in general for southern tropical Mexico, northwest coastal US, and west coastal Canada (high to medium confidence, 26.3.2.1). Many adaptation options currently available can address water supply deficits; adaptation responses to flooding and water quality concerns are more limited (medium confidence) [26.3.3]. Effects of temperature and climate variability on yields of major crops have been observed (high confidence) [25.5.1]. Projected increases in temperature, reductions in precipitation in some regions, and increased frequency of extreme events would result in net productivity declines in major North American crops by the end of the 21st Century without adaptation, although the rate of decline varies by model and scenario, and some regions, particularly in the north, may benefit (very high confidence) [26.5.2]. Given that North America is a significant source of global food supplies, projected productivity declines here may affect global food security (medium confidence). At 2°C, adaptation has high potential to off-set projected declines in yields for many crops, and many strategies offer mitigation co-benefits; but effectiveness of adaptation would be reduced at 4°C (high confidence). [26.5.3] Adaptation capacity varies widely among producers, and institutional support currently lacking in some regions greatly enhances adaptive potential (medium confidence) [26.5.4]. Human health impacts from extreme climate events have been observed, although climate change-related trends and attribution have not been confirmed to-date. Extreme heat events currently result in increases in mortality and morbidity in North America (very high confidence), with impacts that vary by age, location and socioeconomic factors (high confidence) [26.6.1.2]. Extreme coastal storm events can cause excess mortality and morbidity, particularly along the east coast of the United States, and the gulf coast of both Mexico and the United States (high confidence) [26.6.1.1]. A range of water-, food-, and vector-borne infectious diseases, air pollutants, and Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 airborne pollens are influenced by climate variability and change (medium confidence) [26.6.1.3, 26.6.1.4, 26.6.1.5, 26.6.1.6]. Further climate warming in NA will impose stresses on the health sector through more severe extreme events such as heat waves and coastal storms, as well as more gradual changes in climate and CO2 levels. [26.6.2] Human health impacts in NA from future climate extremes can be reduced by adaptation measures such as targeted and sustainable air conditioning, more effective warning and response systems, enhanced pollution controls, urban planning strategies, and resilient health infrastructure (high confidence) [26.6.3]. Observed impacts on livelihoods, economic activities, infrastructure and access to services in North American urban and rural settlements have been attributed to sea level rise, changes in temperature and precipitation, and occurrences of such extreme events as heat waves, droughts and storms (high confidence) [26.8.2.1]. Differences in the severity of climate impacts on human settlements are strongly influenced by context-specific social and environmental factors and processes that contribute to risk, vulnerability and adaptive capacity such as hazard magnitude, populations access to assets, built environment features and governance (high confidence) [26.8.2.1 and 26.8.2.2]. Some of these processes (e.g., the legacy of previous and current stresses) are common to urban and rural settlements, while others are more pertinent to some types of settlements than others. For example, human and capital risks are highly concentrated in some highly exposed urban locations, while in rural areas, geographic isolation and institutional deficits are key sources of vulnerability. Among the most vulnerable are indigenous peoples due to their complex relationship with their ancestral lands and higher reliance on subsistence economies, and those urban centers where high concentrations of populations and economic activities in risk-prone areas combine with several socio-economic and environmental sources of vulnerability (high confidence) [26.8.2.1 and 26.8.2.2]. Although larger urban centers would have higher adaptation capacities, future climate risks from heat waves, droughts, storms and sea level rise in cities would be enhanced by high population density, inadequate infrastructures, lack of institutional capacity and degraded natural environments (high agreement, medium evidence) [26.8.3]. Much of North American infrastructure is currently vulnerable to extreme weather events and, unless investments are made to strengthen them, would be more vulnerable to climate change (medium confidence). Water resources and transportation infrastructure are in many cases deteriorating, thus more vulnerable to extremes than strengthened ones (high confidence). Extreme events have caused significant damage to infrastructure in many parts of North America; risks to infrastructure are particularly acute in Mexico but are a big concern in all three countries (high confidence) [26.7]. Most sectors of the North American economy have been affected by and have responded to extreme weather, including hurricanes, flooding, and intense rainfall (high confidence) (Figure 26-2). Despite a growing experience with reactive adaptation, there are few examples of proactive adaptation anticipating future climate change impacts, and these are largely found in sectors with longer-term decision-making, including energy and public infrastructure. Knowledge about lessons learned and best adaptive practices by industry sector are not well- documented in the published literature [26.7]. There is an emerging concern that dislocation in one sector of the economy may have an adverse impact on other sectors due to supply chain interdependency (medium confidence) [26.7]. Slow onset perils like sea level rise, drought, and permafrost thaw are an emerging concern for some sectors, with large regional variation in awareness and adaptive capacity (medium confidence). Adaptation Responses Adaptation including through technological innovation, institutional strengthening, economic diversification, and infrastructure design can help to reduce risks in the current climate, and to manage future risks in the face of climate change (medium confidence) [26.8.4; 26.9.2]. There is increasing attention to adaptation among planners at all levels of government but particularly at the municipal level, with many jurisdictions engaging in assessment and planning processes. These efforts have revealed the significant challenges and sources of resistance facing planners at both the planning and implementation stages, particularly the adequacy of informational, institutional, financial and human resources, and lack of political will (medium confidence) [26.8.4.2; 26.9.3]. Specific strategies introduced into policy to date tend to be incremental rather than transformational. Fiscal constraints are higher for Mexican jurisdictions and sectors than for Canada or the US. The Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 literature on sectoral-level adaptation is stronger in the areas of technological and engineering adaptation strategies than in social, behavioral and institutional strategies. Adaptation actions have the potential to result in synergies or tradeoffs with mitigation and other development actions and goals (high confidence) [26.8.4.2; 26.9.3]. 26.1. Introduction This chapter assesses literature on observed and projected impacts, vulnerabilities and risks as well as on adaptation practices and options in three North American countries: Canada, Mexico and United States (US). The North American Arctic region is assessed in Chapter 28: Polar Regions. North America ranges from the tropics to frozen tundra, and contains a diversity of topography, ecosystems, economies, governance structures and cultures. As a result, risk and vulnerability to climate variability and change differ considerably across the continent depending on geography, scale, hazard, socio-ecological systems, ecosystems, demographic sectors, cultural values and institutional settings. This chapter seeks to take account of this diversity and complexity as it affects and is projected to affect vulnerabilities, impacts, risks and adaptation across North America. No single chapter would be adequate to cover the range and scope of the literature about climate change vulnerabilities, impacts and adaptations in our three focus countries. (Interested readers are encouraged to review the following reports: (Instituto Nacional de Ecología y Cambio Climático, 2012a; National Climate Assessment Development Advisory Committee, 2013). We therefore attempt to take a more integrative and innovative approach. In addition to describing current and future climatic and socioeconomic trends of relevance to understanding risk and vulnerability in North America (section 26.2), we contrast climate impacts, vulnerabilities and adaptations across and within the three countries in the following key sectors: water resources and management (section 26.3); ecosystems and biodiversity (section 26.4); agriculture and food security (section 26.5); human health (section 26.6); and key economic sectors and services (section 26.7). We use a comparative and place-based approach to explore the factors and processes associated with differences and commonalities in vulnerability, risk and adaptation between urban and rural settlements (section 26.8); and to illustrate and contrast the nuanced challenges and opportunities adaption entails at the city, the subnational and the national level (sections 26.8.4 and 26.9; Box 26-3). We highlight two case studies that cut across sectors, systems or national boundaries. The first, on wildfires (Box 26-2), explores some of the connections between climatic, physical and socioeconomic process (e.g., decadal climatic oscillation, droughts, wildfires land-use, and forest management) and across systems and sectors (e.g., fires direct and indirect impacts on local economies, livelihoods, built environments and human health). The second takes a look at one of the world s longest border between a high-income (US) and middle income country (Mexico) and briefly reflects on the challenges and opportunities of responding to climate change in a transboundary context (Box 26-1). We close with a section (26.10) summarizing key multi-sectoral risks and uncertainties and discussing some of the knowledge gaps that will need to be filled by future research. Findings from the Fourth Assessment Report This section summarizes key findings on North America, as identified in Chapter 13 of the Fourth IPCC assessment focused on Mexico (Magrin et al., 2007), and Chapter 14 on Canada and the US (Field et al., 2007). It focuses on observed and projected impacts, vulnerabilities and risks as well as on adaptation practices and options and highlights areas of agreement and difference between the AR4 s two chapters and our consolidated North American chapter. Observed impacts and processes associated with vulnerability. Both chapter 14 [14.2] and our chapter (Figure 26-2) find that over the past decades, economic damage from severe weather has increased dramatically. Our chapter confirms that although Canada and the US have considerably more adaptive capacity than Mexico, their vulnerability depends on the effectiveness and timing of adaptation and the distribution of capacity, which vary geographically and between sectors. [14.2.6; 14.4; 14.5] [26.2.2; 26.8.2] Chapters 13 and 14 did not assess impacts, vulnerabilities and risks in urban and rural settlements, but rather assessed literature on future risks in the following sectors: Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ecosystems: Both AR4 and our chapter find that ecosystems are under increased stress from increased temperatures, climate variability and other climate stresses (e.g., sea level rise and storm-surge flooding), and that these stresses interact with developmental and environmental stresses (e.g., as salt intrusion, pollution, population growth and the rising value of infrastructure in coastal areas) [13.4.4; 14.2.3; 14.4.3]. Differential capacities for range shifts and constraints from development, habitat fragmentation, invasive species, and broken ecological connections would alter ecosystem structure, function and services in terrestrial ecosystems [14.2; 14.4]. Both reports show that dry soils and warm temperatures are associated with increased wildfire activity and insect outbreaks in Canada and the US [14.2; 14.4; 26.4.2.1]. Water resources: AR4 projects millions in Mexico to be at risk from the lack of adequate water supplies due to climate change[13.4.3]; our chapter, however, finds that water resources are already stressed by non- climatic factors, such as population pressure that will be compounded by climate change [26.3.1]. Both reports find that in the US and Canada rising temperatures would diminish snowpack and increase evaporation [26.2.2.1], thus affecting seasonal availability of water [14.2.1; 26.3.1]. The reports also agree that these effects will be amplified by water demand from economic development, agriculture and population growth, thus imposing further constraints to over-allocated water resources and increasing competition among agricultural, municipal, industrial and ecological uses [14.4.1; 14.4.6; 26.3.3]. Both agree water quality will be further stressed [14.4.1; 26.3.2.2; 13.4.3]. There is more information available now on water adaptation than in AR4 [14.5.1; 26.3.3; 13.5.1.3], and is possible to attribute changes in extreme precipitation, snowmelt and snowpack to climate change [26.3.1; 14.2.1; 13.2.4] Agriculture: The AR4 noted that while increases in grain yields in the US and Canada are projected by most scenarios [14.4.4], in Mexico the picture is mixed for wheat and maize, with different projected impacts depending on scenario used [13.4.2]. Research since the AR4 has offered more cautious projections of yield change in North America due to shifts in temperature and precipitation, particularly by 2100; and significant harvest losses due to recent extreme weather events have been observed [26.5.1]. Furthermore, our chapter reports on recent research that underscores the context specific nature of adaptation capacity and of institutional support and shows that these factors, which greatly enhance adaptive potential, are currently lacking in some regions [26.5.3]. Health: AR4 focused primarily on a set of future health risks. These include changes in the geographical distribution and transmission of diseases such as dengue [13.4.5]; increases in respiratory illness, including exposure to pollen and ozone [14.4] and in mortality from hot temperatures and extreme weather in Canada and the US. AR4 also projects that climate change impacts on infrastructure and human health in cities of Canada and the US would be compounded by aging infrastructure, maladapted urban form and building stock, urban heat islands, air pollution, population growth and an aging population [14.4; 14.5]. Without increased investments in measures such as early warning and surveillance systems, air conditioning, and access to health care, hot temperatures and extreme weather in Canada and the US are predicted to result in increased adverse health impacts [14.4; 14.5]. Our chapter provides a more detailed assessment of these future risks [26.6], besides assessing a richer literature on observed health impacts [26.6.1]. Adaptation: AR4 found that Mexico has early warning and risk management systems, yet it faces planning and management barriers. In Canada and the US, a decentralized response framework has resulted in adaptation that tends to be reactive, unevenly distributed, and focused on coping with rather than preventing problems [14.5]. Both chapters see mainstreaming climate issues into decision making as key to successful adaptation [14.5; 13.5]. The current chapter provides a summary of the growing empirical literature on emerging opportunities and constraints associated with recent institutional adaptation planning activities since the AR4 [26.3.3; 26.4.3; 236.5.3; 26.6.3; 26.8.4;26.9]. In summary, scholarship on climate change impacts, adaptation and vulnerability has grown considerably since the AR4 in North America, particularly in Canada and the US. It is possible now not only to detect and attribute to anthropogenic climate change some impacts such as changes in extreme precipitation, snowmelt and snowpack, but also to examine trends showing increased insect outbreaks, wildfire events and coastal flooding. These latter trends have been shown to be sensitive to climate, but, like the local climate patterns that cause them, have not yet been positively attributed to anthropogenic climate change (see Figure 26-1). Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE 26-1 HERE Figure 26-1 Detection and attribution of climate change impacts. Comparisons of the adequacy of currently available data to detect trends and the degree of understanding of causes of those changes in climatic extreme events in the United States (left; Peterson et al., 2013) and degree of understanding of the climate influence in key impacts in North America (right). Note that climate influence means that the impact has been documented to be sensitive to climate, not that it has been attributed to climate change. Filled boxes indicate that formal detection and attribution to climate change has been performed for the given impact; shaded boxes indicate that a trend has been detected from background variability in the given impact, but formal attribution to climate change has not occurred and the trend could be due to other drivers; and open boxes indicate that a trend has not currently been detected. Key impacts are: 1) earlier peak flow of snowmelt run-off in snow-dominated streams and rivers in western North America [26.3.1], 2) declines in the amount of water stored in spring snowpack in snow-dominated areas of western North America [26.3.1], 3) northward and upward shifts in species distributions in multiple taxa of terrestrial species, although not all taxa and regions [26.4.1], 4) increases in coastal flooding [26.8.1], 5) increases in wildfire activity, including fire season length and area burned by wildfires in the western United States and boreal Canada [Box 26-2], 6) storm-related disaster losses in the United States (most of the increase in insurance claims paid has been attributed to increasing exposure of people and assets in areas of risk) [26.7.6.1, 26.8.1], 7) increases in bark beetle infestation levels in pine tree species in western North America [26.4.2.1], 8) yield increases due in part to increasing temperatures in Canada and higher precipitation in the US; yield variances attributed to climate variability in Ontario and Quebec; yield losses attributed to climate-related extremes across North America [26.5.1], 9) changes in storm-related mortality in the United States [26.6.1.2], 10) changes in heat-related mortality in the United States [26.6.1.2], 11) increases in tree mortality rates in old-growth forests in the western United States and western Canada from 1960-2007 [26.4.2.1], 12) changes in flooding in some urban areas due to extreme rainfall [26.3.1, 26.8.2.1], 13) increase in water supply shortages due to drought [26.8.1, 26.3], and 14) changes in cold- related heat mortality [26.6.1.2].] 26.2. Key Trends Influencing Risk, Vulnerability, and Capacities for Adaptation 26.2.1. Demographic and Socioeconomic Trends 26.2.1.1. Current Trends Canada, Mexico and US share commonalities but also differ in key dimensions shaping risk, vulnerability and adaptation such as population dynamics, economic development, and institutional capacity. During the last years, the three countries, particularly the US, have suffered economic losses from extreme weather events (Figure 26-2). Hurricanes, droughts, floods and other climate-related hazards produce risk as they interact with increases in exposed populations, infrastructure and other assets and with the dynamics of such factors shaping vulnerability as wealth, population size and structure, and poverty (Figure 26-2 and Figure SPM.1). Population growth has been slower in Canada and US than in Mexico (Population Division, Department of Economic and Social Affairs, 2011). Yet population growth in Mexico also decreased from 3.4 percent between 1970-1980 to 1.5 percent yearly during 2000-2010. Populations in the three countries are aging at different rates (Figure 26-2). In 2010, 14.1% of the population in Canada was 60 years and older, compared to 12.7% in the US, and 6.1% in Mexico (Population Division, Department of Economic and Social Affairs, 2011). Urban populations have grown faster than rural populations, resulting in a North America that is highly urbanized (Canada 84.8%, Mexico 82.8% and US 85.8%). Urban populations are also expanding into peri-urban spaces, producing rapid changes in population and land use dynamics that can exacerbate risks from such hazards as floods and wildfires (Eakin et al., 2010; Romero-Lankao et al., 2012a). Mexico has a markedly higher poverty rate (34.8%) than Canada (9.1%) and the US (12.5%) (Figure 26- 2), with weather events and climate affecting poor people s livelihood assets, including crop yields, homes, food security, and sense of place (Chapter 13, section 26.8.2). Between 1970 and 2012, a 10 percent increase in single person households who can be vulnerable because of isolation and low income and housing quality (Roorda et al., 2010), has been detected in the US (Vespa et al., 2013). Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE 26-2 HERE Figure 26-2: Extreme events illustrating vulnerabilities for Mexico, the United States, and Canada. This figure offers a graphic illustration of location of extreme events and relevant vulnerability trends. The observed extreme events have not been attributed to anthropogenic climate change, yet they are climate-sensitive sources of impact illustrating vulnerability of exposed systems, particularly if projected future increases in the frequency and/or intensity of such events should materialize. The figure includes: a) A map (bottom) with population density at 1km resolution highlighting exposure and represented using 2011 Landscan data (Bright et al., 2012). b) A map (top) with significant weather events taking place during 1993-2012. The map only includes disasters with overall losses of more than $1 billion US dollars in US, or more than $500 million US dollars in Mexico and Canada, adjusted to 2012 values (Source: (NatCatSERVICE, 2010). Hence, it does not include the occurrence of disasters of small and medium impact, and it does not capture the impacts of disasters on populations livelihoods and wellbeing. Disasters represented by points that are located at the approximate geographic center of affected regions, frequently span more than one subnational jurisdiction (e.g., the 2012 drought affected 12 Mexican states, Annex Table). c) Four panels (right) with trends in socio-demographic indicators used in the literature to measure vulnerability to hazards (Romero-Lankao et al., 2012): poverty rates, percentage of elderly, GDP per capita and total population (Sources: Comisión Económica para América Latina y el Caribe; U.S. Census Bureau, 2011; Statistics Canada, 2012).] While concentrations of growing populations, water, sanitation, transportation and energy infrastructure and industrial and service sectors in urban areas can be a source of risk, geographic isolation and high dispersion of rural populations also introduce risk because of long distances to essential services (section 26.8.2). Rural populations are more vulnerable to climate events due to smaller labor markets, lower income levels and reduced access to public services. Rural poverty could also be aggravated by changes in agricultural productivity, particularly in Mexico where 65% of the rural population is poor, agricultural income is seasonal, and most households lack insurance (Scott, 2007). Food price increases, which may also result from climate events, would contribute to food insecurity (World Bank, 2011; Lobell et al., 2011). Migration is a key trend affecting North America, recently with movements between urban centers and from rural Mexico into Mexico s cities, and in the US. Rates of migration from rural Mexico are positively associated with natural disaster occurrence and increased poverty trends (Saldana-Zorilla and Sandberg, 2009), and with decreasing precipitation (Nawrotski et al., 2013). Studies of migration induced by past climate variability and change indicate a preference for short-range domestic movement, a complex relationship to assets with indications that the poorest are less able to migrate, and the role of pre-existing immigrant networks in facilitating international migration (Oppenheimer, 2013). North America has become more economically integrated following the 1994 North American Free Trade Agreement. Prior to a 2007-2008 reduction in trade, the three countries registered dynamic growth in industry, employment and global trade of agricultural and manufactured goods (World Bank, 2009). Notwithstanding North America s economic dynamism, increased socioeconomic disparities (Autor et al., 2008) have affected such determinants of vulnerability as differentiated human development and institutional capacity within and across countries. _____ START BOX 26-1 HERE _____ Box 26-1. Adapting in a Transboundary Context: the Mexico-U.S. Border Region Extending over 3111 km (1933 miles; (U.S. Census Bureau, 2011), the border between the United States and Mexico, which can be defined in different ways (Varady and Ward, 2009), illustrates the challenges and opportunities of responding to climate change in a transboundary context. Changing regional climate conditions and socioeconomic processes combined shape differentiated vulnerabilities of exposed populations, infrastructure and economic activities. Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Since at least 1999, the region has experienced high temperatures and aridity anomalies leading to drought conditions (Woodhouse et al., 2010; Wilder et al., 2013) affecting large areas on both sides of the border, and considered the most extreme in over a century of recorded precipitation patterns for the area (Seager and Vecchi, 2010; Cayan et al., 2010; Nielsen-Gammon, 2011). Streamflow in already oversubscribed rivers such as the Colorado and Rio Grande (Nakaegawa et al., 2013) has decreased. Climatological conditions for the area have been unprecedented, with sustained high temperatures that may have exceeded any experienced for 1,200 years. While these changes cannot conclusively be attributed to anthropogenic climate change, they are consistent with climate change projections (Woodhouse et al., 2010). The population of the Mexico-US Border is rapidly growing and urbanizing, doubling from just under 7 million in 1983 to over 15 million in 2012 (Peach and Williams, 2007). Since 1994, rapid growth in the area has been fueled by rapid economic development subsequent to passage of the North American Free Trade Agreement (NAFTA). Between 1990 and 2001 the number of assembly factories or maquiladoras in Mexico grew from 1700 to nearly 3,800, with 2,700 in the border area. By 2004, it was estimated that more than one million Mexicans were employed in more than 3,000 maquiladoras located along the border (U.S. Environmental Protection Agency and Secretaría de Medio Ambiente y Recursos Naturales, 2011; U.S. Environmental Protection Agency, 2012). Notwithstanding this growth, challenges to adaptive capacity include high rates of poverty in a landscape of uneven economic development (Wilder et al., 2013). Large sections of the urban population, particularly in Mexico, live in informal housing lacking the health and safety standards needed to respond to hazards, and with no insurance (Collins et al., 2011). Any effort to increase regional capacity to respond to climate needs to take existing gaps into account. Additionally, there is a prevalence of incipient or actual conflict (Mumme, 1999), given by currently or historically contested allocation of land and water resources (e.g., an over-allocated Colorado river ending in Mexico above the Sea de Cortes (Getches, 2003). Climate change, therefore, would bring additional significant consequences for the region s water resources, ecosystems, and rural and urban settlements. The impacts of regional climatic and non-climatic stresses compound existing urban vulnerabilities that are different across countries. For instance, besides degrading highly diverse ecosystems (Wilder et al., 2013), residential growth in flood-prone areas in Ciudad Juarez has not been complemented with the provision of determinants of adaptive capacity to residents, such as housing, health care and drainage infrastructure. As a result, while differences in mean hazard scores are not significantly different between Ciudad Juarez (Mexico) and El Paso (US), social vulnerability and average risk are three times and two times higher in Ciudad Juárez than in El Paso respectively (Collins, 2008). Projected warming and drying would impose additional burdens on already stressed water resources and ecosystems and compound existing vulnerabilities for populations, infrastructure and economic activities (Wilder et al., 2013). The recent drought in the region illustrated the multiple dimensions of climate-related events, including notable negative impacts on the agricultural sector, water supplies, food security, and risk of wildfire (discussed in Box 26.2) (Wehner et al., 2011; Schwalm et al., 2012; Hoerling et al., 2012). Adaptation opportunities and constraints are shared across international borders, creating the need for cooperation among local, national and international actors. Although there are examples of efforts to manage trans-border environmental issues, such as the US-Mexico International Boundary and Water Commission agreement (International Boundary and Water Commission, 2012), constraints to effective cooperation and collaboration include different governance structures (centralized in Mexico, decentralized in the US); institutional fragmentation; asymmetries in the use and dissemination of information, and language (Wilder et al., 2010; Megdal and Scott, 2011; Wilder et al., 2013). _____ END BOX 26-1 HERE _____ 26.2.1.2. Future Trends The North American population is projected to continue growing, reaching between 531.8 (B2) and 660.1 (A2) million by 2050 (International Institute for Applied System Analysis, 2007). The percentage of elderly people (over Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 64 years) is also projected to continue to increase, by 23.4%-26.9% in Canada, 12.4%-18.4 % in Mexico, and 17.3%-20.9% in the US by 2050 (B2 and A2 respectively) (International Institute for Applied System Analysis, 2007). The elderly are highly vulnerable to extreme weather events (heat waves in particular, Figure 26-2) (Martiello and Giacchi, 2010; Diffenbaugh and Scherer, 2011; Romero-Lankao, 2012; White-Newsome et al., 2012). Numbers of single-person households and female-headed households both of which are vulnerable because of low income and housing quality are anticipated to increase (Roorda et al., 2010). Institutional capacity to address the demands posed by increasing numbers of vulnerable populations may also be limited, with resulting stress on health and the economy. Three other shifts are projected to influence impacts, vulnerabilities and adaptation to climate change in North America: urbanization, migration, and socioeconomic disparity. With small differences between countries, both the concentration of growing populations in some urban areas and the dispersion of rural populations are projected to continue to define North America by 2050. Assuming no change in climate, between 2005 and 2030 the population of Mexico-City-Metro-Area will increase by 17.5%, while between 2007 and 2030 available water will diminish by 11.2% (Romero-Lankao, 2010). Conversely, education, a key determinant of adaptive capacity (Chapter 13), is expected to expand to low-income households, minorities, and women, which could increase the coping capacity of households and have a positive impact on economic growth (Goujon et al., 2004). However, the continuation of current patterns of economic disparity and poverty would hinder future adaptive capacity. Inequality in Mexico is larger (Figure 26-2), having a Gini coefficient (according to which the higher the number the higher economic disparity) of 0.56, in contrast to 0.317 for Canada and 0.389 for the US (Organisation for Economic Co-operation and Development, 2010). Mexico is one of five countries in the world that is projected to experience the highest increases in poverty due to climate-induced extreme events (52% increase in rural households; 95.4% in urban wage-labor households) (CMIP3, A2) (Ahmed et al., 2009). Some studies project increased North American migration in response to climate change. Feng, Krueger and Oppenheimer (2010) estimated the emigration of an additional 1.4 to 6.7 million Mexicans by 2080 based on projected maize yield declines, range depending on model (B1, UKMO and GDFL). Oppenheimer speculates that the indirect impacts of migration could be as substantial as the direct effects of climate change in the receiving area, because the arrival migrants can increase pressure on climate sensitive urban regions (Oppenheimer, 2013, 442). 26.2.2. Physical Climate Trends Some processes important for climate change in North America are assessed in other Chapters of AR5, including WGI Chapter 2 (Observations: Atmosphere and Surface), WGI Chapter 4 (Observations: Cryosphere), WGI Chapter 12 (Long-term Climate Change: Projections, Commitments and Irreversibility), WGI Chapter 14 (Climate Phenomena and their Relevance for Future Regional Climate Change), WGI Annex I (Atlas of Global and Regional Climate Projections), and WGII Chapter 21 (Regional Context). In addition, comparisons of emissions, concentrations, and radiative forcing in the RCPs and SRES scenarios can be found in WGI Annex II (Climate System Scenario Tables). 26.2.2.1. Current Trends It is very likely that mean annual temperature has increased over the past century over most of North America (WGI SPM.1) (Figure 26-3). Observations also show increases in the occurrence of severe hot events over the US over the late 20th century (Kunkel et al., 2008), a result in agreement with observed late-20th-century increases in extremely hot seasons over a region encompassing northern Mexico, the US and parts of eastern Canada (Diffenbaugh and Scherer, 2011). These increases in hot extremes have been accompanied by observed decreases in frost days over much of North America (Alexander et al., 2006; Brown et al., 2010) WGI 2.6.1, decreases in cold spells over the US (Kunkel et al., 2008) WGI 2.6.1, and increasing ratio of record high to low daily temperatures over the US (Meehl et al., 2009). However, warming has been less pronounced and less robust over areas of the central and southeastern US (e.g., (Alexander et al., 2006; Peterson et al., 2008); WGI 2.6.1; WGI SPM.1) (Figure 26-3). It is possible that Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 this pattern of muted temperature change has been influenced by changes in the hydrologic cycle (e.g., (Pan et al., 2004; Portmann et al., 2009), as well as by decadal-scale variability in the ocean (e.g., (Meehl et al., 2012; Kumar et al., 2013b). [INSERT FIGURE 26-3 HERE Figure 26-3: Observed and projected Changes in annual temperature and precipitation. (Top panel, left) observed temperature trends from 1901-2012 determined by linear regression [WGI AR5 Figures SPM.1 and 2.21]. (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression. [WGI AR5 Figure SPM.2] For observed temperature and precipitation, trends have been calculated where sufficient data permits a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level. Diagonal lines indicate areas where change is not significant (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and >90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in >66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5. [Boxes 21-3 and CC-RC]] It is very likely that annual precipitation has increased over the past century over areas of the eastern US and Pacific Northwest (WGI Fig. 2.29) (Figure 26-3). Observations also show increases in heavy precipitation over Mexico, the US and Canada between the mid-20th century and the early 21st century (Peterson and Baringer, 2009; DeGaetano, 2009; Pryor et al., 2009); WGI 2.6.2. Observational analyses of changes in drought are more equivocal over North America, with mixed sign of trend in dryness over Mexico, the US and Canada (WGI 2.6.2 and Fig 2.42) (Dai, 2011; Sheffield et al., 2012). There is also evidence for earlier occurrence of peak flow in snow-dominated rivers globally (Rosenzweig, 2007); WGI 2.6.2). Observed snowpack and snow-dominated runoff have been extensively studied in the western US and western Canada, with observations showing primarily decreasing trends in the amount of water stored in spring snowpack from 1960-2002 (with the most prominent exception being the central and southern Sierra Nevada) (Mote, 2006) and primarily earlier trends in the timing of peak runoff over the 1948-2000 period (Stewart et al., 2006) (WGI 4.5 and Fig. 4.21). Observations also show decreasing mass and length of glaciers in North America (WGI 4.3 and Fig. 4.9, 4.10, 4.11). Further, in assessing changes in the hydrology of the western US, it has been concluded that up to 60% of the climate-related trends of river flow, winter air temperature, and snow pack between 1950 and 1999 are human-induced (Barnett et al., 2008). Observational limitations prohibit conclusions about trends in severe thunderstorms (WGI 2.6.2) and tropical cyclones (WGI 2.6.3) over North America. The most robust trends in extratropical cyclones over North America are determined to be towards more frequent and intense storms over the northern Canadian Arctic and towards less frequent and weaker storms over the southeastern and southwestern coasts of Canada over the 1953-2002 period (WGI 2.7.4)(Wang et al., 2006). WGI concludes that Global mean sea level (GMSL) has risen by 0.19 [0.17 0.21] m over the period 1901 2010 and that it is very likely that the mean rate was 1.7 [1.5 to 1.9] mm yr 1 between 1901 and 2010 and increased to 3.2 [2.8 to 3.6] mm yr 1 between 1993 and 2010 (WGI 3 Executive Summary). In addition, observed changes in extreme sea level have been caused primarily by increases in mean sea level (WGI 3.7.5). Regional variations in the observed rate of sea level rise can result from processes related to atmosphere and ocean variability (such as lower rates along the west coast of the US) or vertical land motion (such as high rates along the US Gulf Coast), but the persistence of the observed regional patterns is unknown (WGI 3.7.3). Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.2.2.2. Climate Change Projections Chapters 11 and 12 of the WGI contribution to the AR5 assess near-term and long-term future climate change, respectively. Chapter 14 of the WGI contribution assesses processes that are important for regional climate change, with section 14.8.3 focused on North America. Many of the WGI conclusions are drawn from Annex I of the WGI contribution to the AR5. The CMIP5 ensemble projects very likely increases in mean annual temperature over North America, with very likely increases in temperature over all land areas in the mid- and late-21st-century periods in RCP2.6 and RCP8.5 (Figure 26-3). Ensemble-mean changes in mean annual temperature exceed 2C over most land areas of all three countries in the mid-21st-century period in RCP8.5 and the late-21st-century period in RCP8.5, and exceed 4C over most land areas of all three countries in the late-21st-century period in RCP8.5. However, ensemble-mean changes in mean annual temperature remain within 2C above the late 20th century baseline over most North American land areas in both the mid- and late-21st-century periods in RCP2.6. The largest changes in mean annual temperature occur over the high latitudes of the United States and Canada, as well as much of eastern Canada, including greater than 6C in the late-21st-century period in RCP8.5. The smallest changes in mean annual temperature occur over areas of southern Mexico, the Pacific Coast of the United States, and the southeastern United States. The CMIP5 ensemble projects warming in all seasons over North America beginning as early as the 2016-2035 period in RCP2.6, with the greatest warming occurring in winter over the high latitudes (WGI Annex I and Figure 26-3)(Diffenbaugh and Giorgi, 2012). The CMIP5 and CMIP3 ensembles suggest that the response of warm-season temperatures to elevated radiative forcing is larger as a fraction of the baseline variability than the response of cold- season temperatures (Diffenbaugh and Scherer, 2011; Kumar et al., 2013b), and the CMIP3 ensemble suggests that the response of temperature in low-latitude areas of North America is larger as a fraction of the baseline variability than the response of temperature in high-latitude areas (Diffenbaugh and Scherer, 2011). In addition, CMIP3 and a high-resolution climate model ensemble suggest that the signal-to-noise ratio of 21st century warming is far greater over the western US, northern Mexico and the northeastern US than over the central and southeastern US (Diffenbaugh et al., 2011), a result that is similar to the observed pattern of temperature trend significance in the US (Figure 26-3). Most land areas north of 45N exhibit likely or very likely increases in mean annual precipitation in the late-21st- century period in RCP8.5 (Figure 26-3). The high latitude areas of North America exhibit very likely changes in mean annual precipitation throughout the illustrative RCP periods, with very likely increases occurring in the mid- 21st-century period in RCP2.6 and becoming generally more widespread at higher levels of forcing. In contrast, much of Mexico exhibits likely decreases in mean annual precipitation beginning in the mid-21st-century period in RCP8.5, with the area of likely decreases expanding to cover most of Mexico and parts of the southcentral and southwestern US in the late-21st-century period in RCP8.5. Likely changes in mean annual precipitation are much less common at lower levels of forcing. For example, likely changes in mean annual precipitation in the mid- and late-21st-century periods in RCP2.6 are primarily confined to increases over areas of Canada and Alaska, with no areas of Mexico and very few areas of the contiguous US exhibiting differences that exceed the baseline variability in more than 66% of the models. CMIP5 projects increases in winter precipitation over Canada and Alaska, consistent with projections of a poleward shift in the dominant cold-season stormtracks (WGI 14.8.3) (Yin, 2005), extratropical cyclones (Trapp et al., 2009) and areas of moisture convergence (WG1 14.8.3), as well as with projections of a shift towards positive North Atlantic Oscillation (NAO) trends (Hori et al., 2007) WGI 14.8.3). CMIP5 also projects decreases in winter precipitation over the southwestern US and much of Mexico associated with the poleward shift in the dominant stormtracks and the expansion of subtropical arid regions (Seager and Vecchi, 2010); WGI 14.8.3). However, there are uncertainties in hydroclimatic change in western North America associated with the response of the tropical Pacific sea surface temperatures (SSTs) to elevated radiative forcing (particularly given the influence of tropical SSTs on the Pacific North American pattern (PNA) and north Pacific storm tracks) (Cayan et al., 1999; Findell and Delworth, 2010; Seager and Vecchi, 2010); WGI 14.8.3), and not all CMIP5 models simulate the observed recent hydrologic trends in the region (Kumar et al., 2013a) Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 For seasonal-scale extremes, CMIP5 projects substantial increases in the occurrence of extremely hot seasons over North America in early, middle and late 21st century periods in RCP8.5 (Diffenbaugh and Giorgi, 2012) (Figure 26- 4). For example, during the 2046-2065 period in RCP8.5, more than 50% of summers exceed the respective late- 20th-century maximum seasonal temperature value over most of the continent. CMIP3 projects similar increases in extremely hot seasons, including greater than 50% of summers exceeding a mid-20th-century baseline throughout much of North America by the mid-21st-century in the A2 scenario (Duffy and Tebaldi, 2012), and greater than 70% of summers exceeding the highest summer temperature observed on record over much of the western US, southeastern US and southern Mexico by the mid-21st-century in the A2 scenario (Battisti and Naylor, 2009). CMIP5 also projects substantial decreases in snow accumulation over the US and Canada (Diffenbaugh et al., 2012) (Figure 26-4), suggesting that the increases in cold-season precipitation over these regions reflect a shift towards increasing fraction of precipitation falling as rain rather than snow (Diffenbaugh et al., 2012). Over much of the western US and western Canada, greater than 80% of years exhibit March snow amount that that is less than the late-20th-century median value beginning in the mid-21st-century period in RCP8.5, with the ensemble-mean change exceeding 2 standard deviations of the ensemble spread. Likewise, greater than 60% of years exhibit March snow amount that is less than the late-20th-century minimum value in the late-21st-century period in RCP8.5, with the ensemble-mean change exceeding 2 standard deviations of the ensemble spread (Diffenbaugh and Giorgi, 2012) (Figure 26-4). CMIP5 also projects increases in the occurrence of extremely dry summer seasons over much of Mexico, the US and southern Canada (Figure 26-4). The largest increases occur over southern Mexico, where greater than 30% of summers in the late-21st-century period in RCP8.5 exhibit seasonal precipitation that is less than the late 20th century minimum summer precipitation. [INSERT FIGURE 26-4 HERE Figure 26-4: Projected changes in extremes in North America. (a) The percentage of years in the 2046 2065 period of RCP8.5 in which the summer temperature is greater than the respective maximum summer temperature of the 1986 2005 baseline period (Diffenbaugh and Giorgi, 2012). (b) The percentage of years in the 2080-2099 period of RCP8.5 in which the summer precipitation is less than the respective minimum summer precipitation of the 1986- 2005 baseline period (Diffenbaugh and Giorgi, 2012) (c) The percentage difference in the 20-year return value of annual precipitation extremes between the 2046-2065 period of RCP4.5 and the 1986-2005 baseline period (from (Kharin et al., 2013). The hatching indicates areas where the differences are not significant at the 5% level. (d) The percentage of years in the 2070-2099 period of RCP8.5 in which the March snow water equivalent is less than the respective minimum March snow water equivalent of the 1976 2005 period (Diffenbaugh et al., 2012). The black (white) stippling indicate areas where the multimodel mean exceeds 1.0 (2.0) standard deviations of the multi-model spread. (a-d) The RCPs and time periods are those used in the peer-reviewed studies in which the panels appear. The 2046-2065 period of RCP8.5 and the 2046-2065 period of RCP4.5 exhibit global warming in the range of 2-3C above the pre-industrial baseline (WGI Fig. 12.40). The 2080-2099 and 2070-2099 periods of RCP8.5 exhibit global warming in the range of 4-5C above the pre-industrial baseline (WGI Fig. 12.40).] For daily-scale extremes, almost all areas of North America exhibit very likely increases of at least 5C in the warmest daily maximum temperature by the late-21st-century period in RCP8.5. Likewise, most areas of Canada exhibit very likely increases of at least 10C in the coldest daily minimum temperature by the late-21st-century period in RCP8.5, while most areas of the US exhibit very likely increases of at least 5C and most areas of Mexico exhibit very likely increases of at least 3C (Sillmann et al., 2013) (WGI Fig. 12.13). In addition, almost all areas of North America exhibit very likely increases of 5% to 20% in the 20-year return value of extreme precipitation by the mid- 21st-century period in RCP4.5 (Figure 26-4), while most areas of the US and Canada exhibit very likely increases of at least 5% in the maximum 5-day precipitation by the late-21st-century period in RCP8.5 (Sillmann et al., 2013)(WGI Fig. 12.13). Further, almost all areas of Mexico exhibit very likely increases in the annual maximum number of consecutive dry days by the late-21st-century period in RCP8.5 (Sillmann et al., 2013) (WGI Fig. 12.13). 26.3. Water Resources and Management Water withdrawals are exceeding stressful levels in many regions of North America such as the southwest US, northern and central Mexico (particularly Mexico City), southern Ontario and the southern Canadian Prairies (Romero-Lankao, 2010; National Water Commission of Mexico, 2010; Sosa-Rodriguez, 2010; Averyt et al., 2011; Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Environment Canada, 2013a). Water quality is also a concern with 10% to 30% of the surface monitoring sites in Mexico having polluted water (National Water Commission of Mexico, 2010), and about 44% of assessed stream miles, and 64% of assessed lake areas in the US not clean enough to support their uses (U.S. Environmental Protection Agency, 2004). Stations in Canada s 16 most populated drainage basins reported at least fair quality, with many reporting good or excellent quality (Environment Canada, 2013b). In basins outside of the populated areas there are some cases of declining water quality where impacts are related to resource extraction, agriculture, and forestry (Hebben, 2009). Water management infrastructure in most areas of North America is in need of repair, replacement or expansion (Section 26.7). Climate change, land use changes and population growth, and demand increases will add to these stresses(U.S. Global Change Research Program, 2009). 26.3.1. Observed Impacts of Climate Change on Water Resources Droughts and Floods: As reported in WG1, Chapter 10 and 26.2.2.1, it is not possible to attribute changes in drought frequency in North America to anthropogenic climate change (Prieto-González et al., 2011; Axelson et al., 2012; Orlowsky and Senevirantne, 2013) (Figure 26-1). Few discernible trends in flooding have been observed in the US (Chapter 3). Changes in the magnitude or frequency of flood events have not been attributed to climate change. Floods are generated by multiple mechanisms (e.g., land use, seasonal changes and urbanization); trend detection is confounded by flow regulation, teleconnections and long-term persistence (section 26.2.2.1; (Kumar et al., 2009; Collins, 2009; Smith et al., 2010; Villarini and Smith, 2010; Villarini et al., 2011; Hirsch and Ryberg, 2012; Prokoph et al., 2012; Instituto Nacional de Ecología y Cambio Climático, 2012a; Peterson et al., 2013). Mean Annual Streamflow: While annual precipitation and runoff increases have been found in the Midwestern and Northwestern United States, decreases have been observed in southern states (Georgakakos et al., 2013) . Chapter 3, WG2 notes the correlation between changes in streamflow and observed regional changes in temperature and precipitation. Kumar et al. (2009) suggest that human activities that have influenced observed trends in streamflow making attribution of changes to climate difficult in many watersheds. Nonetheless, earlier peak flow of snowmelt run-off in snow-dominated streams and rivers in western North America has been formally detected and attributed to anthropogenic climate change (Barnett et al., 2008; Das et al., 2011) (Figure 26-1). Snow Melt: Warm winters produced earlier runoff and discharge but less snow water equivalent and shortened snowmelt seasons in many snow-dominated areas of North America (Barnett et al., 2005; Rood et al., 2008; Reba et al., 2011) (Section 26.2.2., Chapter 3). 26.3.2. Projected Climate Change Impacts and Risks 26.3.2.1. Water Supply Most of this assessment focuses on surface water as there are few groundwater studies (Tremblay et al., 2011; Georgakakos et al., 2013). Impacts and risks vary by region and model used. In arid and semi-arid western US and Canada and in most of Mexico, except the southern tropical area, water supplies are projected to be further stressed by climate change, resulting in less water availability and increased drought conditions (Seager et al., 2007; Instituto Mexicano de Tecnología del Agua, 2010; Cayan et al., 2010; Montero Martinez et al., 2010; MacDonald, 2010; Comisión Nacional del Agua, 2011; Prieto-González et al., 2011; Bonsal et al., 2012; Sosa-Rodriguez, 2013; Orlowsky and Senevirantne, 2013; Diffenbaugh and Field, 2013). Compounding factors include salt water intrusion, and increased groundwater and surface water pollution (Leal Asencio et al., 2008). In the US southwest and southeast, ecosystems and irrigation are projected to be particularly stressed by decreases in water availability due to the combination of climate change, growing water demand, and water transfers to urban Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and industrial users (Seager et al., 2009; Georgakakos et al., 2013). In the Colorado River Basin crop irrigation requirements for pasture grass are projected to increase by 20% by 2040 and by 31% by 2070 (Dwyer et al., 2012). In the Rio Grande basin, New Mexico, runoff is projected decrease by 8%-30% by 2080 due to climate change. Water transfers may entail significant transaction costs associated with adjudication and potential litigation; and might have economic, environmental, social, and cultural impacts that vary by water user (Hurd and Coonrod, 2012). In Mexico, water shortages combined with increased water demands are projected to increase surface and groundwater over exploitation (Comisión Nacional del Agua, 2011). Other parts of North American are projected to have different climate risks. The vulnerability of water resources over the tropical southern region of Mexico is projected to be low for 2050: precipitation decreases from 10%-5% in the summer and no precipitation changes in the winter. After 2050, greater winter precipitation is projected, increasing the possibility of damaging hydropower and water storage damns by floods, while precipitation is projected to decrease by 40%-35% in the summer (Instituto Mexicano de Tecnología del Agua, 2010). Throughout the 21st century, cities in NW Washington are projected to have drawdown of average seasonal reservoir storage in the absence of demand reduction because of less snow pack even though annual stream flows increase. Without accounting for demanding increases, projected reliability of all systems remains above 98% through mid and late century (Vano et al., 2010a; Comisión Nacional del Agua, 2011). Throughout the eastern US, water supply systems will be negatively impacted by lost snowpack storage; rising sea levels contributing to increased storm intensities and salt water intrusion; possibly lower stream-flows; land use and population changes; and other stresses (Sun et al., 2008; Obeysekera et al., 2011). In Canada s Pacific Northwest Region, cool season flows are expected to increase, while warm seasons flows would decrease (Hamlet, 2011). Southern Alberta, where approximately two-thirds of Canadian irrigated land is located, is projected to experience declines in mean annual stream flow, especially during the summer (Shepherd et al., 2010; Poirier and de Loë, 2012; Tanzeeba and Gan, 2012). In the Athabasca River Basin in northern Alberta, modeling results consistently indicate large projected declines in mean annual flows (Kerkhoven and Gan, 2011). In contrast, modeling results for basins in Manitoba indicate an increase in mean annual runoff (Choi et al., 2009). Some model results for the Fraser River Basin in British Columbia indicate increases in mean annual runoff by the end of the 21st Century, while others indicate decreases (Kerkhoven and Gan, 2011). In central Quebec, (Chen et al., 2011b) project a general increase in discharge during November-April, and a general decrease in summer discharge under most climate change conditions. 26.3.2.2. Water Quality Many recent studies project water quality declines due to the combined impacts of climate change and development (Daley et al., 2009; Tu, 2009; Praskievicz and Chang, 2011; Wilson and Weng, 2011; Tong et al., 2012). Increased wildfires linked to a warming climate are expected to affect water quality downstream of forested headwater regions (Emelko et al., 2011). Model simulation of lakes under a range of plausible higher air temperatures (Tahoe, Great Lakes, Lake Onondaga and shallow polymictic lakes), depending on the system, predict a range of impacts such as increased phytoplankton, fish and cyanobacteria biomass, lengthened stratification periods with risks of significant hypolimnetic oxygen deficits in late summer with solubilization of accumulated phosphorous and heavy metals with accelerated reaction rates, and decreased lake clarity (Dupuis and Hann, 2009; Trumpickas et al., 2009; Sahoo et al., 2011; Taner et al., 2011). Model simulations have found seasonal climate change impacts on nonpoint source pollution loads, while others have found no impact (Marshall and Randhir, 2008; Tu, 2009; Taner et al., 2011; Praskievicz and Chang, 2011). Changes in physical-chemical-biological parameters and micropollutants are predicted to negatively affect drinking water treatment and distribution systems (Delpla et al., 2009; Carriere et al., 2010; Emelko et al., 2011). Wastewater treatment plants would be more vulnerable as increases in rainfall and wet weather lead to higher rates of inflow and infiltration (New York City Department of Environmental Protection, 2008; King County Department of Natural Resources and Parks, 2008; Flood and Cahoon, 2011). They would also face reduced hydraulic capacities due to higher sea levels and increased river and coastal flooding (Flood and Cahoon, 2011), with higher sea levels also Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 threatening sewage collection systems (Rosenzweig et al., 2007; King County Department of Natural Resources and Parks, 2008) 26.3.2.3. Flooding Projected increases in flooding (Georgakakos et al., 2013) may affect sectors ranging from agriculture and livestock in southern tropical Mexico (National Water Commission of Mexico, 2010) to urban and water infrastructure in areas such as Dayton, Ohio, metro Boston and the Californian Bay-Delta region (Committee on Flood Control Alternatives in the American River Basin et al., 1995; Kirshen et al., 2006; California Department of Water Resources, 2009; Wu, 2010). Floods could begin earlier, have earlier peaks and longer durations (e.g., southern Quebec basin). Urbanization can compound the impacts of increased flooding due to climate change, particularly in the absence of flood management infrastructure that takes climate change into account (Hejazi and Markus, 2009; Mailhot and Duchesne, 2010; Sosa-Rodriguez, 2010). (Ntelekos et al., 2010) estimate that annual riverine flood losses in the USA could increase from approximately $2 billion now to $7-$19 billion annually by 2100 depending upon emission scenario and economic growth rate. 26.3.2.4 Instream Uses Projections of climate impacts on instream uses vary by region and time-frame. Hydropower generation, affected by reduced lake levels, is projected to decrease in arid and semi-arid areas of Mexico (Comisión Intersecretarial de Cambio Climatico, 2009; Sosa-Rodriguez, 2013) and in the Great Lakes (Buttle et al., 2004; Mortsch et al., 2006; Georgakakos et al., 2013). In the US Pacific Northwest under several emission scenarios. it is projected to increase in 2040 by approximately 5% in the winter and decrease by approximately 13% in the summer, with annual reductions of approximately 2.5%. Larger increases and decreases are projected by 2080 (Hamlet et al., 2010). On the Peribonka River system in Quebec, annual mean hydropower production will similarly decrease in the short- term increase by as much as 18 % in late 21st century (Minville et al., 2009). Navigation on the Great Lakes, Mississippi River and other inland waterways may benefit from less ice cover but will be hindered by increased floods and low river levels during droughts (Georgakakos et al., 2013). 26.3.3. Adaptation There are a range of structural and non-structural adaptation measures being implemented with many being no- regret policies. For instance, in preparation for more intense storms, New York City is using green infrastructure to capture rainwater before it can flood the combined sewer system and is elevating boilers and other equipment above ground (Bloomberg, 2012). The Mexican cities of Monterrey, Guadalajara, Mexico City and Tlaxcala are reducing leaks from water systems (Comisión Intersecretarial de Cambio Climatico, 2009; National Water Commission of Mexico, 2010; Sosa-Rodriguez, 2010; Romero-Lankao, 2010). Regina, SK has increased urban water conservation efforts (Natural Resources Canada, 2008). The 540-foot high, 1300-foot long concrete Ross Dam in the state of Washington, US was built on a special foundation so it could later be raised in height (Simmons, 1974). Dock owners in the Trent-Severn Waterway in the Great Lakes have moved their docks into deeper water to better manage impacts on shorelines (Coleman et al., 2013). The South Florida Water Management District is assessing the vulnerability to sea level rise of its aging coastal flood control system and exploring adaptation strategies, including a strategy known as forward pumping (Obeysekera et al., 2011). In Cambridge, Ontario, extra capacity culverts are being installed in anticipation of larger runoff (Scheckenberger et al., 2009). Water meters have been installed to reduce consumption by different users such as Mexican and Canadian farmers and in households of several Canadian cities (Instituto Nacional de Ecología y Cambio Climático, 2006; Natural Resources Canada, 2008). Agreements and regulations are underway such as the 2009 SECURE Water Act which establishes a federal climate change adaptation program with required studies to assess future water supply risks in Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 the western U.S (42 USC § 10363). One such large, multi-year study was recently completed in the US for the Colorado River (Bureau of Reclamation, 2013) and others are planned. Agreements and regulations are underway, such as the 2007 Shortage Sharing Agreement for the management of the Colorado River, driven by concerns about water conservation, planning, better reservoir coordination, and preserving flexibility to respond to climate change(Bureau of Reclamation, 2007). Quebec Province is requiring dam safety inspections every 10 years to account for new knowledge on climate change impacts (Centre d'expertise hydrique du Québec, 2003). Expanded beyond flood and hydropower management to now include climate change, the Columbia River Treaty is a good example of an international treaty to manage a range of water resources challenges(U.S. Army Corps of Engineers and Bonneville Power Administration, 2013). 26.4. Ecosystems and Biodiversity 26.4.1. Overview Recent research has documented gradual changes in physiology, phenology and distributions in North American ecosystems consistent with warming trends (Dumais and Prévost, 2007). Changes in phenology and species distributions, particularly in the United States and Canada, have been attributed to rising temperatures, which have in turn been attributed to anthropogenic climate change via joint attribution (Root et al., 2005; Vose et al., 2012). Concomitant with 20th century temperature increases, northward and upward shifts in plant, mammal, bird, lizard, and insect species distributions have been documented extensively in the western United States and eastern Mexico (Parmesan, 2006; Kelly and Goulden, 2008; Moritz et al., 2009; Tingley et al., 2009; Sinervo et al., 2010). These distribution shifts consistent with climate-change interact with other environmental changes such as land-use change, hindering the ability of species to respond (Ponce-Reyes et al., 2013). A range of techniques have been applied to assess the vulnerability of North American ecosystems and species to changes in climate (Loarie et al., 2009; Anderson et al., 2009; Glick and Stein, 2011). A global risk analysis based on dynamic global vegetation models identified boreal forest in Canada as notably vulnerable to ecosystem shift (Scholze et al., 2006). Since the AR4, the role of extreme events, including droughts, flood, hurricanes, storm surges, and heat waves, is a more prominent theme in studies of climate change impacts on North American ecosystems (Chambers et al., 2007; IPCC, 2012). A number of ecosystems in North America are vulnerable to climate change. For example, species in alpine ecosystems are at high risk due to limited geographic space into which to expand (Villers-Ruiz and Castaneda- Aguado, In press). Many forest ecosystems are susceptible to wildfire and large-scale mortality and infestations events (section 26.4.1). Across the continent, potentially rapid rates of climate change may require location shifts at velocities well outside the range in historical reconstructions (Sandel et al., 2011; Schloss et al., 2012). Changes in temperature, precipitation amount, and carbon dioxide concentrations can have different effects across species and ecological communities (Parmesan, 2006; Matthews et al., 2011), leading to ecosystem disruption and reorganization (Smith et al., 2011; Dukes et al., 2011), as well as movement or loss. The following section focuses in more depth on climate vulnerabilities in forests and coastal ecosystems. These ecosystems span all three North American countries, are illustrative cases of where understanding the opportunities for conservation and adaptation practices is important, and recent research advances and new evidence of increased vulnerabilities since AR4 motivate further exploration. Further treatment of grasslands and shrublands can be found in AR5 WGII 4.3.3.2.2, wetlands and peatlands 4.3.3.3, and tundra, alpine, and permafrost systems in 4.3.3.4. Additional synthesis of climate change impacts on terrestrial, coastal and ocean ecosystems can be found in Chapter 8 of the US National Climate Assessment (Groffman et al., 2013). Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.4.2. Tree Mortality and Forest Infestation 26.4.2.1. Observed Impacts Droughts of unusual severity, extent, and duration have affected large parts of western and southwestern North America and resulted in regional-scale forest dieback in Canada, US and Mexico. Extensive tree mortality has been related to drought exacerbated by high summertime temperatures in trembling aspen (Populus tremuloides), pinyon pine (Pinus edulis) and lodgepole pine (Pinus contorta) since the early 2000s (Breshears et al., 2005; Hogg et al., 2008; Raffa et al., 2008; Michaelian et al., 2011; Anderegg et al., 2012). In 2011 and 2012 forest dieback in Northern and central Mexico was associated with extreme temperatures and severe droughts (Comisión Nacional Forestal, 2012a).Widespread forest-mortality events triggered by extreme climate events can alter ecosystem structure and function (Phillips et al., 2009; Allen et al., 2010; Anderegg et al., 2013). Similarly, multi-decadal changes in demographic rates, particularly mortality, indicate climate-mediated changes in forest communities over longer periods (Hogg and Bernier, 2005; Williamson et al., 2009). Average annual mortality rates increased from less than 0.5% of trees per year in the 1960s in forests of western Canada and the US to, respectively, 1.5-2.5% (Peng et al., 2011), and 1.0-1.5% in the 2000s in the US (van Mantgem et al., 2009). The influences of climate change on ecosystem disturbance, such as insect outbreaks have become increasingly salient and suggest that these disturbances could have a major influence on North American ecosystems and economy in a changing climate. In terms of carbon stores these outbreaks have the potential to turn forests into carbon sources (Kurz et al., 2008a; Kurz et al., 2008b; Hicke et al., 2012). Warm winters in western Canada and US have increased winter survival of the larvae of bark beetles, helping drive large-scale forest infestations and forest die-off in western North America since the early 2000s (Bentz et al., 2010). Beginning in 1994, mountain pine beetle outbreaks have severely affected over 18 million hectares of pine forests in British Columbia, and outbreaks are expanding northwards (Energy, Mines and Resources: Forest Management Branch, 2012). 26.4.2.2. Projected Impacts and Risks Projected increases in drought severity in southwestern forests and woodlands in United States and northwestern Mexico suggest that these ecosystems may be increasingly vulnerable, with impacts including vegetation mortality (Seager and Vecchi, 2010; Williams et al., 2010; Overpeck and Udall, 2010) and an increase of biological agents such as beetles, borers, pathogenic fungi, budworms and other pests (Drake et al., 2005). An index of forest drought stress calibrated from tree rings indicates that projected drought stress by the 2050s in the SRES A2 scenario from the CMIP3 model ensemble, due primarily to warming-induced rises in vapor pressure deficit, exceeds the most severe droughts of the past 1,000 years (Park et al., 2013). Under a scenario with large changes in global temperature (SRES A2) increases in growing-season temperature in forest soils in southern Quebec are as high as 5.0 C towards the end of the century and decreases of soil water content reach 20-40% due to elevated evapotranspiration rates (Houle et al., 2012). More frequent droughts in tropical forests may change forest structure and regional distribution, favoring a higher prevalence of deciduous species in the forests of Mexico (Drake et al., 2005; Trejo et al., 2011). Shifts in climate are expected to lead to changes in forest infestation, including shifts of insect and pathogen distributions into higher latitudes and elevations (Bentz et al., 2010). Predicted climate warming is expected to have effects on bark beetle population dynamics in the western United States, western Canada, and northern Mexico that may include increases in developmental rates, generations per year, and changes in habitat suitability (Waring et al., 2009). As a result, the impacts of bark beetles on forest resources are expected to increase (Waring et al., 2009). Wildfire, a potentially powerful influence on North American forests in the 21st century, is discussed in Box 26-2. Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.4.3. Coastal Ecosystems Highly productive estuaries, coastal marshes and mangrove ecosystems are present along the Gulf coast and the East and West coasts of North America. These ecosystems are subject to a wide range of non-climate stressors, including urban and tourist developments and the indirect effects of overfishing (Mortsch et al., 2006; Bhatti et al., 2006; Lund et al., 2007; Comisión Nacional para el Conocimiento y Uso de la Biodiversidad et al., 2007). Climate change adds risks from sea-level rise, warming, ocean acidification, extratropical cyclones, altered upwelling, and hurricanes and other storms. 26.4.3.1. Observed Climate Impacts and Vulnerabilities Sea level rise, which has not been uniform across the coasts of North America (Crawford et al., 2007; Kemp et al., 2008; Leonard et al., 2009; Zavala-Hidalgo et al., 2010; Sallenger et al., 2012), is directly related to flooding and loss of coastal dunes and wetlands, oyster beds, seagrass and mangroves (Feagin et al., 2005; Cooper et al., 2008; Najjar et al., 2010; Ruggiero et al., 2010; McKee, 2011; Martinez Arroyo et al., 2011). Increases in sea surface temperature in estuaries alter metabolism, threatening species, especially cold water fish (Crawford et al., 2007). Historical warm periods have coincided with low salmon abundance and restriction of fisheries in Alaska (Crozier et al., 2008; U.S. Global Change Research Program, 2009). North Atlantic cetaceans, and tropical coral reefs in the Gulf of California and the Caribbean have been affected by increases in the incidence of diseases associated with warm waters and low water quality (Mumby et al., 2011; International Council for the Exploration of the Sea, 2011). Increased concentrations of CO2 in the atmosphere due to human emissions are causing ocean acidification (Executive summary chapters 5 and 6; FAQ 5.1). Along the temperate coasts of North America acidification directly affects calcareous organisms, including colonial mussel beds, with indirect influences on food webs of benthic species (Wootton et al., 2008). Increased acidity in conjunction with high temperatures has been identified as a serious threat to coral reefs and other marine ecosystems in the Bahamas and the Gulf of California(Doney et al., 2009; Hernández et al., 2010; Mumby et al., 2011). Tropical storms and hurricanes can have a wide range of effects on coastal ecosystems, potentially altering hydrology, geomorphology (erosion), biotic structure in reefs and nutrient cycling. Hurricane impacts on the coastline change dramatically the marine habitat of sea turtles, reducing feeding habitats, such as coral reefs and areas of seaweed, and nesting places. (Márquez, R. and Jiménez, Ma. del C., 2010; Liceaga-Correa et al., 2011) 26.4.3.2. Projected Impacts and Risks Projected increases in sea levels, particularly along the coastlines of Florida, Louisiana, North Carolina, and Texas (Kemp et al., 2008; Leonard et al., 2009; Weiss et al., 2011), will threaten many plants in coastal ecosystems through increased inundation, erosion, and salinity levels. In settings where landward shifts are not possible, a 1 m rise in sea level will result in loss of wetlands and mangroves along the Gulf of Mexico of 20% (in Tamaulipas) to 94% (in Veracruz) (Flores Verdugo et al., 2010). Projected impacts of increased water temperatures include contraction of coldwater fish habitat and expansion of warm-water fish habitat (Mantua et al., 2010), which can increase the presence of invasive species that threaten resident populations (Janetos et al., 2008). Depending on scenario, Chinook salmon in the Pacific Northwest may decline by 20 to 50% by 2040-50 (Battin et al., 2007; Crozier et al., 2008), integrating across restrictions in productivity and abundance at the southern end of their range and expansions at the northern end (Azumaya et al., 2007), although habitat restoration and protection particularly at lower elevations may help mitigate declines in abundance. Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Continuing ocean acidification will decrease coral growth and interactions with temperature increases will lead to increased risk of coral bleaching, leading to declines in coral ecosystem biodiversity (Veron et al., 2009) (5.4.2.4, Box CC-OA). Oyster larvae in Chesapeake Bay grew more slowly when reared with CO2 levels between 560 and 480ppm compared to current environmental conditions (Gazeau et al., 2007; Miller et al., 2009; Najjar et al., 2010). While future trends in thunderstorms and tropical cyclones are uncertain (26.2.2), any changes, particularly an increase in the frequency of category 4 and 5 storms (Bender et al., 2010; Knutson et al., 2010) could have profound impacts on mangrove ecosystems, which require 25 years for recovery from storm damage (Kovacs et al., 2004; Flores Verdugo et al., 2010). 26.4.4. Ecosystems Adaptation, and Mitigation In North America, a number of adaptation strategies, are being applied in novel and flexible ways to address the impacts of climate change (Mawdsley et al., 2009; National Oceanic and Atmospheric Administration, 2010; Gleeson et al., 2011; Poiani et al., 2011). The best of these are based on detailed knowledge of the vulnerabilities and sensitivities of species and ecosystems, and with a focus on opportunities for building resilience through effective ecosystem management. Government agencies and nonprofit organizations have established initiatives that emphasize the value of collaborative dialogue between scientists and practitioners, indigenous communities, and grass-roots organizations to develop no-regrets and co-benefits adaptation strategies (Ogden and Innes, 2009; Gleeson et al., 2011; Halofsky et al., 2011; Cross et al., 2012; Instituto Nacional de Ecología y Cambio Climático, 2012b; Cross et al., 2013). Examples of adaptation measures implemented to respond to climate change impacts on ecosystems are diverse. They include programs to reduce the incidence of Canadian forest pest infestations (Johnston et al., 2010); breeding programs for resistance to diseases and insect pests (Yanchuk and Allard, 2009); use of forest programs to reduce the incidence of forest fires and encourage agroforestry in areas of Mexico (Sosa-Rodriguez, 2013); and selection by forest or fisheries managers of activities that are more adapted to new climatic conditions (Vasseur and Catto, 2008). Example programs have addressed commercial fishing, mass tourism (Pratchett et al., 2008), and enforcement mechanisms for using water regulation technologies to maintain quantity and quality in wetlands around the Great Lakes and San Francisco, California (Mortsch et al., 2006; Okey et al., 2012). Assisted migration is increasingly discussed as a potential management option to maintain health and productivity of forests; yet the technique has logistical and feasibility challenges (Keel, 2007; Hoegh-Guldberg et al., 2008; Winder et al., 2011). Several lines of evidence indicate that effective adaptation requires changes in approach and becomes much more difficult if warming exceeds 2°C above preindustrial levels (Comisión Nacional para el Conocimiento y Uso de la Biodiversidad et al., 2007; Mansourian et al., 2009; U.S. Forest Service, 2010; Glick and Stein, 2011; Barragan et al., 2011; Instituto Nacional de Ecología y Cambio Climático, 2012b). Even though options for effective adaptation are increasingly constrained at warming over 2°C, some opportunities will remain. In particular, efforts to maintain or increase forest carbon stocks can lead to numerous benefits, including not only benefits for atmospheric CO2 (Anderson and Bell, 2009; Anderson et al., 2011). Even where there are opportunities, managers face challenges in designing management practices that favor carbon stocks, while at the same time maintaining biodiversity, recognizing the rights of indigenous people, and contributing to local economic development (Food and Agriculture Organization, 2012). _____ START BOX 26-2 HERE _____ Box 26-2. Wildfires Wildfire is a natural process, critical to nutrient cycling, controlling populations of pests and pathogens, biodiversity and fire-adapted species (Bond and Van Wilgen, 1996). However, since the mid-1980s large wildfire activity in North America has been marked by increased frequency and duration, and longer wildfire seasons (Westerling et al., 2006; Williamson et al., 2009). Recent wildfires in western Canada, the US and Mexico relate to long and warm spring and summer droughts, particularly when they are accompanied by winds (Holden et al., 2007; Comisión Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Nacional Forestal, 2012b). Interacting processes such as land-use changes associated with the expansion of settlements and activities in peri-urban areas or forested areas, combined with the legacies of historic forest management that prescribed fire suppression, also substantially increase wildfire risk (Radeloff et al., 2005; Peter et al., 2006; Theobald and Romme, 2007; Fischlin et al., 2007; Gude et al., 2008; Collins and Bolin, 2009; Hammer et al., 2009; Brenkert-Smith, 2010). Drought conditions are strongly associated with wildfire occurrence, as dead fuels such as needles and dried stems promote the incidence of firebrands and spot fires (Keeley and Zedler, 2009; Liu et al., 2012). Drought trends vary across regions (Groisman et al., 2007; Girardin et al., 2012): the western US has experienced drier conditions since the 1970s (Peterson et al., 2013); drought periods in Alberta and Idaho have coincided with large burned areas (Pierce and Meyer, 2008; Kulshreshtha, 2011); heterogeneous patterns of drought severity and a reduction of wildfire risk have been detected for the circumboreal region (Girardin et al., 2009). Decadal climatic oscillations also contribute to differences in drought, and thus in wildfire occurrences. The areas burned in the continent boreal forest and in northwest and central Mexico correlate to the dynamics of seasonal land/ocean temperature variability (Macias Fauria and Johnson, 2006; Skinner et al., 2006; Villers-Ruíz and Hernández-Lozano, 2007; Macias Fauria and Johnson, 2008; Girardin and Sauchyn, 2008); which is shifting toward hotter temperatures and longer droughts. Such human practices as slash-and-burn agriculture can have negative impacts on Mexican forests (Bond and Keeley, 2005; Comisión Nacional de Áreas Naturales Protegidas and The Nature Conservancy, Programa México, 2009). Drought index projections and climate change regional models show increases in wildfire risk during the summer and fall on the southeast Pacific coast, Northern Plains and the Rocky Mountains (Liu et al., 2012). In places like Sierra Nevada, mixed conifer forests, which have a natural cycle of small, non-crown fires, are projected to have massive crown-fires (Bond and Keeley, 2005) (Table 26-1). While healthy forests (Davis, 2004) and many fire-maintained systems that burn at lower intensities can provide carbon sequestration and thus mitigation co-benefits (e.g., Longleaf pine savanna, Sierra mixed-conifer)(Fried et al., 2008; North et al., 2012), forests affected by pests and fires are less effective carbon sinks, and wildfires themselves are a source of emissions. Wildfires pose a direct threat to human lives, property and health. Over the last 30 years, 155 people were killed in wildfires across North America, including 103 in the United States, 50 in Mexico and 2 in Canada (Centre for Research on the Epidemeology of Disasters, 2012). Direct effects include injury and respiratory effects from smoke inhalation, with firefighters at increased risk (Naeher et al., 2007; Reisen and Brown, 2009; Reisen et al., 2011). Wildfire activity causes impacts on human health (section 26.6). Minimizing adverse effects of wildfires involves short-term and long-term strategies such as planned manipulation of vegetation composition and stand structure (Girardin et al., 2012; Terrier et al., 2013), suppression of fires where required, fuel treatments, use of fire-safe materials in construction, community planning, and reduction of arson. Not all negative consequences of fire can be avoided, though a mixture of techniques can be used to minimize adverse effects (Girardin et al., 2012). Prescribed fire may be an important tool for managing fire risk in Canada and the US (Hurteau and North, 2010; Wiedinmyer and Hurteau, 2010; Hurteau et al., 2011). Managers in the US have encouraged reduction of flammable vegetation around structures with different levels of success (Stewart et al., 2006). However, such efforts depend largely on land-use planning, the socio-economic capacity of communities at risk, the extent of resource dependence, community composition, and the risk perceptions, attitudes and beliefs of decision-makers, private property owners, and affected populations (McFarlane, 2006; Repetto, 2008; Collins and Bolin, 2009; Martin et al., 2009; Trainor et al., 2009; Brenkert-Smith, 2010). Indigenous peoples are at higher risk from wildfire and may have unique requirements for adaptation strategies (Carroll et al., 2010; Christianson et al., 2012a; Christianson et al., 2012b). Effective forest management requires stakeholder involvement and investment. The provision of adequate information on smoke, prescribed fire, pest management, and forest thinning is crucial, as is building trust between stakeholders and land managers (Dombeck et al., 2004; Flint et al., 2008; Chang et al., 2009). Institutional shifts Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 from reliance on historical records toward incorporation of climate forecasting in forest management is also crucial to effective adaptation (McKenzie et al., 2004; Millar et al., 2007; Kolden and Brown, 2010). _____ END BOX 26-2 HERE _____ 26.5. Agriculture and Food Security Projected declines in global agricultural productivity (Chapter 7) have implications for food security among North Americans. Because North America is a major exporter (Food and Agriculture Organization, 2009; Schlenker and Roberts, 2009), shifts in agricultural productivity here may have implications for global food security. Canada and the US are relatively food secure, although households living in poverty are vulnerable. 17.6% of Mexicans are food insecure (Monterroso et al., 2012). Indigenous peoples are highly vulnerable due to high reliance on subsistence (Chapter 12). While this section focuses on agricultural production, food security is related to multiple factors (See Chapter 7). 26.5.1. Observed Climate Change Impacts Historic yield increases are attributed in part to increasing temperatures in Canada and higher precipitation in the US (Pearson et al., 2008; Sakurai et al., 2011; Nadler and Bullock, 2011); (high agreement, medium evidence), although multiple non-climatic factors affect historic production rates. In many North American regions optimum temperatures have been reached for dominant crops, thus continued regional warming would diminish rather than enhance yields (Jones et al., 2005) (high confidence). Regional yield variances over time have been attributed to climate variability e.g., Ontario (Cabas et al., 2010) and Quebec (Almaraz et al., 2008). Since 1999 a marked increase in crop losses attributed to climate-related events such as drought, extreme heat and storms has been observed across North America (Hatfield et al., 2013), with significant negative economic effects (Swanson et al., 2007; Chen and McCarl, 2009; Costello et al., 2009) (high confidence). In Mexico, agriculture accounted for 80% of weather-related financial losses since 1990 (Saldana-Zorrilla, 2008) (Figure 26-2). 26.5.2. Projected Climate Change Risks Studies project productivity gains in northern regions and where water is not projected to be a limiting factor, across models, time frames and scenarios (Hatfield et al., 2008; Pearson et al., 2008; Wheaton et al., 2010; Stöckle et al., 2010); (high confidence). Overall yields of major crops in North America are projected to decline modestly by mid- century and more steeply by 2100 among studies that do not consider adaptation (very high confidence). Certain regions and crops may experience gains in the absence of extreme events, and projected yields vary by climate model (Paudel and Hatch, 2012; Liu et al., 2013). Among studies projecting yield declines, two factors stand out: exceedance of temperature thresholds, and water availability. Yields of several important North American agriculture sectors including grains, forage, livestock and dairy decline significantly above temperature thresholds (Wolfe et al., 2008; Schlenker and Roberts, 2009; Craine et al., 2010) Temperature increases affect product quality as well e.g., coffee (Lin, 2007), wine grapes (Hayhoe et al., 2004; Jones et al., 2005), wheat (Porter and Semenov, 2005), fruits and nuts (Lobell et al., 2006), and cattle forage (Craine et al., 2010). Projected temperature increases would reduce corn, soy and cotton yields by 2020, with declines ranging from 30- 82% by 2099 depending on crop and scenario (steepest decline for corn, A1) (Schlenker and Roberts, 2009). Studies also project increasing inter-annual yield variability over time (Sakurai et al., 2011; Urban et al., 2012). Several studies focus on California, one of North America s most productive agricultural regions. Modest and variable yield changes among several California crops are projected to 2026, with yield declines from 9-29% by 2097 (A2, DAYCENT model). (Lobell and Field, 2011; Lee et al., 2011) found little negative effect for California perennials by 2050 due to projected climate change, assuming irrigation access (GCM ensemble, A2 and B1). (Hannah et al., Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2013), however, project large declines in land suitability for California viticulture by 2050 (with increases further north) with RCPs 4.5 and 8.5 (GCM ensemble); declines greater under RCP 8.5. Heat-induced livestock stress, combined with reduced forage quality, would reduce milk production and weight gain in cattle (Wolfe et al., 2008; Hernandez et al., 2011). Precipitation increases off-set but do not entirely compensate for temperature-related declines in productivity (Kucharik and Serbin, 2008). In regions projected to experience increasing temperatures combined with declining precipitation, declines in yield and quality are more acute (Craine et al., 2010; Monterroso Rivas et al., 2011). Projected change in climate will reduce soil moisture and water availability in the US Western/Southwest, the Western Prairies in Canada, and central and northern Mexico (Pearson et al., 2008; U.S. Global Change Research Program, 2009; Cai et al., 2009; Esqueda et al., 2010; Vano et al., 2010b; Kulshreshtha, 2011) (very high confidence). CMIP5 models indicate soil moisture decreases across the continent in Spring and Summer under RCP8.5, with high agreement (Dirmeyer et al., 2013). Based on a combined exposure/consumptive water use model, the US Great Plains is identified as one of four global future vulnerability hotspots for water availability from the 2030s and beyond, where anticipated water withdrawals would exceed 40% of freshwater resources (Liu et al., 2013). In western US and Canada, projected earlier Spring snowmelt and reduced snowpack would affect productivity negatively regardless of precipitation, as water availability in Summer and Fall are reduced (Schlenker et al., 2007; Forbes et al., 2011; Kienzle et al., 2012). Projected increases in extreme heat, drought and storms affect productivity negatively (Chen and McCarl, 2009; Kulshreshtha, 2011). The northeastern and southeastern US have been identified as vulnerability hotspots for corn and wheat production respectively by 2045 with vulnerability worsening thereafter, using a combined drought exposure and adaptive capacity assessment, with only slight differences between A1B and B2 scenarios (Fraser et al., 2013). Central North America is identified as among the globe s regions of highest risk of heat stress by 2070 (NIES GCM; A1B) (Teixeira et al., 2013). 26.5.3. A Closer Look at Mexico Much of Mexico s landbase is already marginal for two of the country s major crops: corn and beef (Buechler, 2009). Severe desertification in Mexico due to non-climate drivers further compromises productivity (Huber- Sannwald et al., 2006). Land classified suitable for rain-fed corn is projected to decrease from 6.2% currently to between 3% and 4.3% by 2050 (UKHadley B2, ECHAM5/MPI A2) (Monterroso Rivas et al., 2011). The distribution of most races of corn is expected to be reduced and some eliminated by 2030 (A2, three climate models) (Ureta et al., 2012). Precipitation declines of 0-30% are projected over Mexico by 2040; with the most acute declines in Northwestern Mexico, the primary region of irrigated grain farming (declines steeper in A2 than A1B (18 model ensemble). Although projected increases in precipitation may contribute to increase in rangeland productivity in some regions (Monterroso Rivas et al., 2011), a study in Veracruz indicates that the effects of projected maximum summer temperatures on livestock heat stress are expected to reach the Danger level (at which losses can occur) by 2020 and continue to rise (A2, B2, three GCMs) (Hernandez et al., 2011). Coffee, an economically important crop supporting 500,000 primarily indigenous households (González Martínez, 2006), is projected to decline 34% by 2020 in Veracruz if historic temperature and precipitation trends continue (Gay et al., 2006); see also (Schroth et al., 2009), on declines in Chiapas). Many of Mexico s agricultural communities are also considered highly vulnerable, due to high sensitivity and/or low adaptive capacity (Monterroso et al., 2012). The agriculture sector here consists primarily of small farmers (Claridades Agropecuarias 2006), who face high livelihood risks due to limited access to credit and insurance (Eakin and Tucker, 2006; Wehbe et al., 2008; Saldana-Zorilla and Sandberg, 2009; Walthall et al., 2012). Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.5.4. Adaptation The North American agricultural industry has the adaptive capacity to off-set projected yield declines and capitalize on opportunities under 2° warming. (Butler and Huybers, 2012) project a reduction in US corn yield loss from 14% to 6% with 2° warming, with spatial shifts in varietal selection (not accounting for variability in temperature and precipitation). Incremental strategies, such as planting varieties better suited to future climate conditions and changing planting dates, have been observed across the continent (Bootsma et al., 2005; Conde et al., 2006; Eakin and Appendini, 2008; Coles and Scott, 2009; Nadler and Bullock, 2011; Paudel and Hatch, 2012; Campos et al., 2013). In some sectors we are seeing multi-organizational investments in adaptation. International coffee retailers and non-governmental organizations, for example, are engaged in enhancing coffee farmers adaptive capacity (Schroth et al., 2009; Soto-Pinto and Anzueto, 2010). Other strategies specifically recommended for Mexico include soil remediation; improved use of climate information; rainwater capture and drip irrigation (Sosa-Rodriguez, 2013). New crop varieties better suited to future climates, including GMOs, are under development in the US (e.g. (Chen et al., 2012), although potential risks have been noted (Quist and Chapela, 2001). Current trends in agricultural practices in commercial regions such as the Midwestern US, however, amplify productivity risks posed by climate change (Hatfield et al., 2013). Incremental strategies will have reduced effectiveness under a 2099/4°C warming scenario, which would require more systemic adaptation, including production and livelihood diversification (Howden et al., 2007; Mehta et al., 2013; Smith and Gregory, 2013; Asseng et al., 2013). Some adaptive strategies impose financial costs and risks onto producers (Craine et al., 2010;Wolfe et al., 2008), which may be beyond the means of smallholders (Mercer et al., 2012) or economically precluded for low-value crops. Technological improvements improve yields under normal conditions but do not protect harvests from extremes (U.S. Global Change Research Program, 2009;Wittrock et al., 2011). Others may have maladaptive effects (e.g. increased groundwater and energy consumption). Crop-specific weather index insurance, for example (widely implemented in Mexico to support small farmers), may impose disincentives to invest in diversification and irrigation (Fuchs and Wolff, 2010). Many strategies have co-benefits, however, in fact investments in agricultural adaptation represent a cost-effective mitigation strategy (Lobell et al., 2013). Low- and no-till practices reduce soil erosion and runoff, protect crops from extreme precipitation (Zhang and Nearing, 2005), retain soil moisture, reduce biogenic and geogenic greenhouse gas emissions (Nelson et al., 2009; Suddick et al., 2010), and build soil organic carbon (Aguilera et al., 2013). Planting legumes and weed management on pastures enhance both forage productivity and soil carbon sequestration (Follett and Reed, 2010). Shade perennials increase soil moisture retention (Lin, 2010) and contribute to local cooling (Georgescu et al., 2011). Crop diversification mediates the impacts of climate and market shocks (Eakin and Appendini, 2008) and enhances management flexibility (Chhetri et al., 2010). Barriers and Enablers Market forces and technical feasibility alone are insufficient to foster sectoral-level adaptation (Kulshreshtha, 2011). Institutional support is key, found to be inadequate in many contexts (Bryant et al., 2008; Klerkx and Leeuwis, 2009; Jacques et al., 2010; Tarnoczi and Berkes, 2010; Brooks and Loevinsohn, 2011; Alam et al., 2012; Anderson and McLachlan, 2012)(high confidence). Even many suggested adaptation strategies with anticipated economic benefits are often not adopted by farmers, suggesting the need for more attention to culture and behavior (Moran et al., 2013). Attitudinal studies among US farmers indicate limited acknowledgement of anthropogenic climate change, associated with lower levels of support for adaptation (Arbuckle Jr et al., 2013; Gramig et al., 2013) (high agreement, medium evidence). Other key enablers are access to and quality of information (Tarnoczi and Berkes, 2010; Tarnoczi, 2011; Baumgart- Getz et al., 2012; Tambo and Abdoulaye, 2012), particularly regarding optimum crop management, production inputs and optimum crop-specific geographic information. Social networks are important for information dissemination and farmer support (Chiffoleau, 2009; Wittrock et al., 2011; Baumgart-Getz et al., 2012). Networks among producers may be especially important to the level of awareness and concern farmers hold about climate Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 change(Frank et al., 2010; Sánchez-Cortés and Chavero, 2011), while also enabling extensive farmer-to-farmer exchange of adaptation strategies (Eakin et al., 2009). 26.6. Human Health Large national assessments of climate and health have been carried out in the US and Canada (Belanger et al., 2008a) (see references in 26.1). These have highlighted the potential for changes in impacts of extreme storm and heat events, air pollution, pollen, and infectious diseases, drawing from a growing NA research base analyzing observed and projected relationships among weather, vulnerability and health. The causal pathways leading from climate to health are complex, and can be modified by factors including economic status, pre-existing illness, age, other health risk factors, access to health care, built and natural environments, adaptation actions and others. Human health is an important dimension of adaptation planning at the local level, much of which has so-far focused on warning and response systems to extreme heat events (New York State Climate Action Council, 2012). 26.6.1. Observed Impacts, Vulnerabilities and Trends 26.6.1.1. Storm-Related Impacts The magnitude of health impacts of extreme storms depends on interactions between exposure and characteristics of the affected communities (Keim, 2008). Coastal and low-lying infrastructure and populations can be vulnerable due to flood-related interruptions in communications, healthcare access, and mobility. Health impacts can arise through direct pathways of traumatic death and injury (e.g., drowning, impacts of blowing and falling objects; contact with power wires) as well as more indirect, longer-term pathways related to damage to health and transportation infrastructure, contamination of water and soil, vector-borne diseases, respiratory diseases and mental health (Gamble et al., 2008). Infectious disease impacts from flooding include creation of breeding sites for vectors (Ivers and Ryan, 2006) and bacterial transmission through contaminated water and food sources causing gastrointestinal disease. Chemical toxins can be mobilized from industrial or contaminated sites (Euripidou and Murray, 2004). Elevated indoor mold levels associated with flooding of buildings and standing water are identified as risk factors for cough, wheeze and childhood asthma (Bornehag et al., 2001; Jaakkola et al., 2005). Mental health impacts can arise due to the stress of evacuation, property damage, economic loss, and household disruption (Weisler et al., 2006; Gamble et al., 2008; Berry et al., 2010; Bethel et al., 2011). Since 1970, there has been no clear trend in US hurricane deaths, once the singular Katrina event is set aside (Blake et al., 2007). 26.6.1.2. Temperature Extremes Studies throughout North America have shown that high temperatures can increase the mortality and/or morbidity (e.g., Medina-Ramon and Schwartz, 2007; Kovats and Hajat, 2008; O'Neill and Ebi, 2009; Anderson and Bell, 2009; Knowlton et al., 2009; Deschenes et al., 2009; Kenny et al., 2010; Hajat and Kosatsky, 2010; Cueva-Luna et al., 2011; Hurtado-Díaz et al., 2011; Romero-Lankao et al., 2012b). Extremely cold temperatures have also been associated with increased mortality (Medina-Ramon and Schwartz, 2007), an effect separate from the seasonal phenomenon of excess winter mortality, which does not appear to be directly related to cold temperatures (Kinney, 2012). To date, trends over time in cold-related deaths have not been investigated. Most available NA evidence derives from the US and Canada, though one study reported significant heat- and cold-related mortality impacts in Mexico City (McMichael et al., 2008). US EPA has tracked the death rate in the US from 1979 to 2009 for which death certificates list the underlying cause of death as heat related (U.S. Environmental Protection Agency, Office of Atmospheric Programs, 2012). No clear trend upwards or downwards is yet apparent in this indicator. Note that this case definition is thought to significantly underestimate the total impacts of heat on mortality. Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.6.1.3. Air Quality Ozone and particulate matter (e.g., PM2.5 and PM10) have been associated with adverse health effects in many locations in North America (Romero-Lankao et al., 2013b). Emissions, transport, dilution, chemical transformation, and eventual deposition of air pollutants all can be influenced by meteorological variables such as temperature, humidity, wind speed and direction, and mixing height (Kinney, 2008). Although air pollution emission trends will play a dominant role in future pollution levels, climate change may make it harder to achieve some air quality goals (Jacob and Winner, 2009). Forest fire is a source of particle emissions in NA, and can lead to increased cardiac and respiratory-disease incidence, as well as direct mortality (Rittmaster et al., 2006; Ebi et al., 2008). The indoor environment also can affect health in many ways, e.g., via penetration of outdoor pollution, emissions or pollutants indoors, moisture-related problems, and transmission of respiratory infections. Indoor moisture leads to mold growth, a problem that is exacerbated in colder regions such as northern NA in the winter (Potera, 2011). Climate variability and change will affect indoor air quality, but with direction and magnitude that remains largely unknown (Institute of Medicine, 2011). 26.6.1.4. Pollen Exposure to pollen has been associated with a range of allergic outcomes, including exacerbations of allergic rhinitis (Cakmak et al., 2002; Villeneuve et al., 2006) and asthma (Delfino, 2002). Temperature and precipitation in the months prior to the pollen season affect production of many types of tree and grass pollen (Reiss and Kostic, 1976; Minero et al., 1998; Lo and Levetin, 2007; U.S. Environmental Protection Agency, 2008). Ragweed pollen production is responsive to temperatures and to CO2 concentrations (Ziska and Caulfield, 2000; Wayne et al., 2002; Ziska et al., 2003; Singer et al., 2005). Because pollen production and release can be affected by temperature, precipitation, and CO2 concentrations, pollen exposure and allergic disease morbidity could change in response to climate change. However, to date, the timing of the pollen season is the only evidence for observed climate-related impacts. Many studies have indicated that pollen seasons are beginning earlier (Emberlin et al., 2002; Rasmussen, 2002; Clot, 2003; Teranishi et al., 2006; Frei and Gassner, 2008; Levetin and Van, 2008; Ariano et al., 2010). Ragweed season length has increased at some monitoring stations in the United States (Ziska et al., 2011). Research on trends in NA has been hampered by the lack of long-term, consistently collected pollen records (U.S. Environmental Protection Agency, 2008). 26.6.1.5. Waterborne Diseases Waterborne infections are an important source of morbidity and mortality in North America. Commonly reported infectious agents in US and Canadian outbreaks include legionella bacterium, the cryptosporidium parasite, campylobacter, and giardia (Belanger et al., 2008b; Center for Disease Control and Prevention, 2011). Cholera remains an important agent in Mexico (Greer et al., 2008). Risk of waterborne illness is greater among the poor, infants, elderly, pregnant women, and immune-compromised individuals (Rose et al., 2001; Gamble et al., 2008) . In Mexico City, declining water quality has led to ineffective disinfection of drinking water supplies (Mazari-Hiriart et al., 2005; Sosa-Rodriguez, 2010). Changes in temperature and hydrological cycles can influence the risk of waterborne diseases (Curriero et al., 2001; Greer et al., 2008; Harper et al., 2011). Severe storms have been shown to play a role in water-borne disease risks in Canada (Thomas et al., 2006). Floods enhance the potential for runoff to carry sediment and pollutants to water supplies (Karl et al., 2008). Disparities in access to treated water were identified as a key determinant of under age-5 morbidity due to water borne illnesses in the central State of Mexico (Jimenez-Moleon and Gomez-Albores, 2011). 26.6.1.6. Vectorborne Diseases The extent to which climate change has altered, and will alter, the geographic distribution of vectors of infectious disease remains uncertain because of the inherent complexity of the ecological system. Spatial and temporal Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 distribution of disease vectors depend not only on climate factors, but also on land use/change, socio-economic and socio-cultural factors, prioritization of vector control, access to health care and human behavioral responses to perception of disease risk, among other factors (Lafferty, 2009; Wilson, 2009). Although temperature drives important biological processes in these organisms, climate variability on a daily, seasonal or interannual scale may result in organism adaptation and shifts, though not necessarily expansion, in geographic range (Lafferty, 2009; Tabachnick, 2010; McGregor, 2011). Range shifts may alter the incidence of disease depending on host receptiveness and immunity, as well as the ability of the pathogen to evolve so that strains are more effectively and efficiently acquired (Reiter, 2008; Beebe et al., 2009; Rosenthal, 2009; Russell, 2009; Epstein, 2010). North Americans are currently at risk from a number of vector-borne diseases, including Lyme disease (Ogden et al., 2008; Diuk-Wasser et al., 2010), dengue fever (Jury, 2008; Ramos et al., 2008; Johansson et al., 2009; Kolivras, 2010; Degallier et al., 2010; Lambrechts et al., 2011; Riojas-Rodriguez et al., 2011; Lozano-Fuentes et al., 2012), West Nile virus (Morin and Comrie, 2010; Gong et al., 2011) and Rocky Mountain spotted fever, to name a few. Risk is increasing from invasive vector-borne pathogens, such as chikungunya (Ruiz-Moreno et al., 2012) and Rift Valley fever viruses (Greer et al., 2008). There is also potential risk from invasive vector-borne pathogens, such as chikungunya (Ruiz-Moreno et al., 2012) and Rift Valley fever viruses (Greer et al., 2008). Mexico is listed as high risk for dengue fever by the WHO. There has been an increasing number of cases of Lyme disease in Canada and Lyme disease vectors are spreading along climate-determined trajectories (Leighton et al., 2012; Koffi et al., 2012). 26.6.2. Projected Climate Change Impacts Projecting future consequences of climate warming for heat-related mortality and morbidity is challenging, due in large part to uncertainties in the nature and pace of adaptations that populations and societal infrastructure will undergo in response to long-term climate change (Kinney et al., 2008). Additional uncertainties arise from changes over time in population demographics, economic well-being, and underlying disease risk, as well as in the model- based predictions of future climate and our understanding of the exposure-response relationship for heat-related mortality. However, climate warming will lead to continuing health stresses related to extreme high temperatures, particularly for the northern parts of North America. The health implications of warming winters remain uncertain (Kinney, 2012). Several recent studies have projected future health impacts due to air pollution in a changing climate (Knowlton et al., 2004; Bell et al., 2007; Tagaris et al., 2009; Tagaris et al., 2010; Chang et al., 2010). There is a large literature examining future climate influences on outdoor air quality in North America, particularly for ozone (Murazaki and Hess, 2006; Steiner et al., 2006; Tao et al., 2007; Kunkel et al., 2007; Holloway et al., 2008; Lin et al., 2008; Nolte et al., 2008; Wu et al., 2008; Avise et al., 2009; Chen et al., 2009; Liao et al., 2009; Racherla and Adams, 2009; Lin et al., 2010; Tai et al., 2010). This work suggests with medium confidence that ozone concentrations could increase under future climate change scenarios if emissions of precursors were held constant (Jacob and Winner, 2009). However, analyses show that future increases can be offset through measures taken to limit emission of pollutants (Kelly et al., 2012). The literature for PM2.5 is more limited than that for ozone, and shows a more complex pattern of climate sensitivities, with no clear net influence of warming temperatures (Liao et al., 2007; Tagaris et al., 2008; Avise et al., 2009; Pye et al., 2009; Mahmud et al., 2010). On the other hand, PM2.5 plays a crucial role. Regarding outdoor pollen, warming will lead to further changes in the seasonal timing of pollen release (high confidence). Another driver of future pollen could be changing spatial patterns of vegetation as a result of climate change. Regarding clean water supplies, extreme precipitation can overwhelm combined sewer systems and lead to overflow events that can threaten human health (Patz et al., 2008). Conditional on a future increase in such events, we can anticipate increasing risks related to water-borne diseases. Whether future warmer winters in the United States and Canada will promote transmission of diseases like dengue and malaria is uncertain, in part, because of access to amenities such as screening and air-conditioning that provide barriers to human-vector contact. Socio-economic factors also play important roles in determining risks. Better longitudinal datasets and empirical models are needed to address research gaps on climate-sensitive infectious diseases, as well as to provide a better mechanism for weighting the roles of external drivers such as climate change Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 on a macro/micro scale, human-environmental changes on a regional to local scale, and extrinsic factors in the transmission of vector-borne infectious diseases (Wilson, 2009; McGregor, 2011). 26.6.3. Adaptation Responses Early warning and response systems can be developed to build resilience to events like heat waves, storms and floods (Ebi, 2011) and protect susceptible populations, which include infants, children, the elderly, individuals with pre-existing diseases, and those living in socially and/or economically disadvantaged conditions (Pinkerton et al., 2012). Adaptation planning at all scales to build resilience for health systems in the face of a changing climate is a growing priority (Kinney et al., 2011). Adaptation to heat events can occur via physiologic mechanisms, indoor climate control, urban-scale cooling initiatives, and with implementation of warning and response systems (Romero-Lankao et al., 2012b). Additional research is needed on the extent to which warning systems prevent deaths (Harlan and Ruddell, 2011). Efforts to reduce GHG emissions could provide health co-benefits, including reductions in heat-related and respiratory illnesses (Health Chapter, US National Climate Assessment). 26.7. Key Economic Sectors and Services There is mounting evidence that many economic sectors across North America have experienced climate impacts and are adapting to the risk of loss and damage from weather perils. This section covers the literature for the energy, transportation, mining, manufacturing, construction and housing, and insurance sectors in North America. Recent studies find a range of adaptive practices and adaptation responses to experience with extreme events, and only an emerging consideration of proactive adaptation in anticipation of future global warming. 26.7.1. Energy 26.7.1.1. Observed Impacts Energy demand for cooling has increased as building stock and air conditioning penetration have increased (Wilbanks et al., 2012). Extreme weather currently poses risk to the energy system (Wilbanks et al., 2012). For example, Hurriacne Sandy results in a loss of power to 8.5 million customers in the Northeast US (National Oceanic and Atmospheric Administration, 2013). Energy consumption is a major user of water resources in North America, with 49% of the water withdrawals in the US for thermoelectric power (Kenny et al., 2009). 26.7.1.2. Projected Impacts Demand for summer cooling is projected to increase and demand for winter heating is projected to decrease. Total energy demand in North America is projected to increase in coming decades because of non-climate factors (Galindo, 2009; National Energy Board, 2011; Energy Information Administration, 2013). Climate change is projected to have varying geographic impacts. In Canada, a net decrease in residential annual energy demand is projected by 2050 and by 2100 (Isaac and Van Vuuren, 2009; Schaeffer et al., 2012). It is difficult to project changes in net energy demand in the US because of uncertainties in such factors as climate change, and change in technology, population, and energy prices. Peak demand for electricity is projected to increase more than the average demand for electricity, with capacity expansion needed in many areas (Wilbanks et al., 2012). Given the projected increases in energy demand in the southern United States from climate change (Auffhammer and Aroonruengsawat, 2011; Auffhammer and Aroonruengsawat, 2012) it is reasonable to conclude that Mexico will have a net increase in demand. Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Major water resource related concerns include effects of increased cooling and other demands for water and water- scarcity in the west; effects of extreme weather events, sea-level rise, hurricanes, and seasonal droughts in the southeast; and effects of increased cooling demands in the northern regions (Wilbanks et al., 2008; Wilbanks et al., 2012; McDonald, 2012; U.S. Department of Energy, 2013) . The magnitude of projected impacts on hydropower potential will vary significantly between regions and within drainage basins(Desrochers et al., 2009; Shrestha et al., 2012; Kienzle et al., 2012). Annual mean hydropower production in the Peribonka River in Quebec is estimated to increase by approximately 10% by mid-century and 20% late in the century under the A2 scenario (Minville et al., 2009). Higher temperatures and increased climate variability can have adverse impacts on renewable energy production such as wind and solar (U.S. Department of Energy, 2013). Changing cloud cover affects solar energy resources, changes in winds affect wind power potentials, and temperature change and water availability can affect biomass production (Wilbanks et al., 2008; U.S. Department of Energy, 2013) . 26.7.1.3. Adaptation Many adaptations are underway to reduce vulnerability of the energy sector to extreme climate events such as heat, drought, and flooding (U.S. Department of Energy, 2013). Adaptation includes many approaches such as increased supply and demand efficiency (e.g., through more use of insulation), more use of urban vegetation and reflective surfaces, improved electric grid, reduced reliance on above ground distribution systems, and distributed power (Wilbanks et al., 2012). Important barriers to adaptation include uncertainty about future climate change, inadequate information on costs of adaptation, lack of climate resilient energy technologies, and limited price signals (U.S. Department of Energy, 2013). Strategies resulting in energy demand reduction would reduce greenhouse gas emissions and reduce the vulnerability of the sector to climate change. 26.7.2. Transportation 26.7.2.1. Observed Impacts Much of the transportation infrastructure across North America is aging, or inadequate (Mexico) which may make it more vulnerable to damage from extreme events and climate change. Approximately 11% of all US bridges are structurally deficient, 20% of airport runways are in fair or poor condition, and more than half of all locks are more than 50 years old (US Department of Transportation 2013). More than US$2 trillion is needed to bring infrastructure in the US up to good condition (American Society of Civil Engineers, 2009) p. 6. Canadian infrastructure had an investment deficit of C$125 billion in the 1980s and 1990s (Mirza and Haider, 2003). Some transportation systems have been harmed (Figure 26-2). For example, in 2008, Hurricane Ike caused $2.4 billion in damages to ports and waterways in Texas. (McDonald, 2012). The superflood in Tennessee and Kentucky in 2010 caused $2.3 billion in damage (National Oceanic and Atmospheric Administration, 2013). Hurricane Sandy resulted in flooding of portions of New York City s subway system, overtopping of runways at La Guardia airport, and caused $400 million in damage to the New Jersey transit system (National Oceanic and Atmospheric Administration, 2013). 26.7.2.2. Projected Impacts Scholarship on projected climate impacts on transportation infrastructure focuses mostly on US and Canada. Increases in high temperatures, intense precipitation, drought, sea level, and storm surge could affect transportation across the United States. The greatest risks would be to coastal transportation infrastructure, but there could be benefits to marine and lake transportation in high latitudes from less ice cover (Transportation Research Board, Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2008). A 1-meter sea level rise combined with a 7-meter storm surge could inundate over half of the highways, arterials, and rail lines in the US Gulf coast (Savonis et al., 2008). Declining water levels in the Great Lakes would increase shipping costs by restricting vessel drafts and reducing vessel cargo volume (Millerd, 2011). In southern Canada by the 2050s, cracking of roads from freeze and thaw would decrease under the B2 and A2 scenarios, structures would freeze later and thaw earlier, while higher extreme temperatures could increase rutting (Mills et al., 2009) and related maintenance and rehabilitation costs (Canadian Council of Professional Engineers, 2008). A 1 to 1.5°C increase in global mean temperature would increase the costs of keeping paved and unpaved roads in the United States in service by, respectively, US$2 to US$3 billion per year by 2050 (Chinowsky et al., 2013). Tens of thousands to more than 100,000 bridges in the US could be vulnerable to increasing peak river flows in the mid- and late-21st Century under the A1B and A2 scenarios. Strengthening vulnerable bridges to be less vulnerable to climate change is estimated to cost approximately US$100 to $250 billion (Wright et al., 2012). 26.7.2.3. Adaptation Adaptation steps are being taken in North America, particularly to protect transportation infrastructure from sea level rise and storm surge in coastal regions. Almost all of the major river and bay bridges destroyed by Hurricane Katrina surge waters were rebuilt at higher elevations, and the design of the connections between the bridge decks and piers were strengthened (Grenzeback and Luckmann, 2006). Adaptation actions include protecting coastal transportation from sea level rise and more intense coastal storms or possibly relocating infrastructure. Many Midwestern states are examining channel protection and drainage designs, while transportation agencies in Canada and the United States have been preparing to manage the aftermath of extreme weather events (Meyer et al., 2013). In addition, new materials may be needed so pavement and rail lines can better withstand more extreme temperatures. 26.7.3. Mining 26.7.3.1. Observed Impacts Climatic sensitivities of mining activities, including exploration, extraction, processing, operations, transportation and site remediation, have been noted in the limited literature (Chiotti and Lavender, 2008; Furgal and Prowse, 2008; Meza-Figueroa et al., 2009; Ford et al., 2010a; Locke et al., 2011; Gómez-álvarez et al., 2011; Kirchner et al., 2011; Pearce et al., 2011; Stratos Inc, 2011). Drought-like conditions have affected the mining sector by limiting water supply for operations (Pearce et al., 2011), enhancing dust emissions from quarries (Pearce et al., 2011) and increasing concentrations of heavy metals in sediments (Gómez-álvarez et al., 2011). Heavy precipitation events have caused untreated mining wastewater to be flushed into river systems (Pearce et al., 2011). High loads of contamination (from metals, sulfate and acid) at three mine sites in the United States were measured during rainstorm events following dry periods (Nordstrom, 2009). 26.7.3.2. Projected Impacts Climate change is perceived by Canadian mine practitioners as an emerging risk, and in some cases, a potential opportunity (Ford et al., 2010a; Ford et al., 2011; Pearce et al., 2011; National Round Table on the Environment and the Economy, 2012), with potential impacts on transportation (Ford et al., 2011) and limited water availability (Acclimatise, 2009) from projected drier conditions (Sun et al., 2008; Seager and Vecchi, 2010) being identified as key issues. An increase in heavy precipitation events projected for much of North America (Warren and Egginton, 2008; Nordstrom, 2009) would adversely affect the mining sector. A study on acid rock damage drainage in Canada concluded that an increase in heavy precipitation events presented a risk of both environmental impacts and Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 economic costs (Stratos Inc, 2011) Damage to mining infrastructure from extreme events, for active and post- operation mines, is also a concern (Pearce et al., 2011). Climate change impacts that affect the bottom-line of mining companies (through direct impacts or associated costs of adaptation), would have consequences for employment, for both the mining sectors and local support industries (Backus et al., 2013). 26.7.3.3. Adaptation Despite increasing awareness, there are presently few documented examples of proactive adaptation planning within the mining sector (Acclimatise, 2009; Ford et al., 2010a; Ford et al., 2011). However, adjustments to management practices to deal with short-term water shortages, including reducing water intake, increasing recycling and establishing infrastructure to move water from tailing ponds, pits and quarries, have worked successfully in the past (Chiotti and Lavender, 2008). Integrating climate change considerations at the mine planning and design phase increases the opportunity for effective and cost-efficient adaptation (Stratos Inc, 2011). 26.7.4. Manufacturing 26.7.4.1. Observed Impacts There is little literature focused on climate change and manufacturing, although one study suggested that manufacturing is among the most sensitive sectors to weather in the US (Lazo et al., 2011). Weather affects the supply of raw material, production process, transportation of goods, and demand for certain products. In 2011, automobile manufacturers in North America experienced production losses associated with shortages of components due to flooding in Thailand (Newswire, 2011). In 2013, reduced cattle-supply and higher feed prices associated with drought in Texas led to a decision to close a beef processing plant (Beef Today Editors, 2013). Drought also caused delays for barge shipping on the Mississippi River in 2012 (Reuters, 2012). Major storms, like Hurricanes Sandy, Katrina and Andrew, significantly disrupted manufacturing activities, including plant shutdowns due to direct damages and/or loss of electricity and supply disruptions due to unavailability of parts, and difficulties delivering products due to compromised transportation networks (Baade et al., 2007; Dolfman et al., 2007). 26.7.4.2. Projected Impacts The drier conditions (Sun et al., 2008; Seager and Vecchi, 2010; Wehner et al., 2011) would present challenges, especially for manufacturers located in regions already experiencing water stress. This could lead to increased conflicts over water between sectors and regions, and affect the ability of regions to attract new facilities or retain existing operations. A study of the effect of changes in precipitation (A1B scenario) on 70 industries in the US between 2010 and 2050 found potentially significant losses in production and employment due to declines in water availability and the interconnectedness of different industries (Backus et al., 2013). Another potential concern for manufacturing relates to impacts of heat on worker safety and productivity. Several studies suggest that higher temperatures and humidity would lead to decreased productivity and increased occupational health risks (e.g., Kjellstrom et al., 2009; Hanna et al., 2011; Kjellstrom and Crowe, 2011). 26.7.4.3. Adaptation Some companies are beginning to recognize the risks climate change presents to their manufacturing operations, and consider strategies to build resilience (National Round Table on the Environment and the Economy, 2012). Coca Cola has a water stewardship strategy focusing on improving water use efficiency at its manufacturing plants, while Rio Tinto Alcan is assessing climate change risks for their operations and infrastructure, which include vulnerability of transport systems, increased maintenance costs, and disruptions due to extreme events (National Round Table on the Environment and the Economy, 2012). Air conditioning is a viable and effective adaptation option to address Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 some of the impacts of warming, though it does incur greater demands for electricity and additional costs (Scott et al., 2008a). Sourcing raw materials from different regions and relocating manufacturing plants are other adaptation strategies that can be used to increase resiliency and reduce vulnerability. 26.7.5. Construction and Housing 26.7.5.1. Observed Impacts The risk of damage from climate change is important for construction industries, though little research has systematically explored the topic (Morton et al., 2011). Private data from insurance companies report a significant increase in severe weather damage to buildings and other insured infrastructure over several decades (Munich Re, 2012). 26.7.5.2. Projected Impacts Most studies project a significant further increase in damage to homes, buildings and infrastructure (Bjarndadottir et al., 2011; IPCC, 2012). Affordable adaptation in design and construction practices could reduce much of the risk of climate damage for new buildings and infrastructure, involving reform in Building Codes and other standards (Feltmate and Thistlethwaite, 2012). However, adaptation best practices in design and construction are often prohibitively expensive to apply to existing buildings and infrastructure, so much of the projected increase in climate damage risk involves existing buildings and infrastructure. 26.7.5.3. Adaptation Engineering and construction knowledge exists to design and construct new buildings to accommodate the risk of damage from historic extremes and anticipated changes in severe weather (Kelly, 2010; Ministry of Municipal Affairs and Housing, 2011; Insurance Institute for Business and Home Safety, 2012). Older buildings may be retrofit to increase resilience, but these changes are often more expensive to introduce into an existing structure than if they were included during initial construction. The housing and construction industries have made advances toward climate change mitigation by incorporating energy efficiency in building design (Heap, 2007). Less progress has been made in addressing the risk of damage from extreme weather events (Kenter, 2010). In some markets, like the Gulf Coast of the United States, change is under way in the design and construction of new homes in reaction to recent hurricanes (Levina et al., 2007; Kunreuther and Michel-Kerjan, 2009; Insurance Institute for Business and Home Safety, 2011), but in most markets across North America there has been little change in building practices. The cost of adaptation measures combined with limited long-term liability for future buildings has influenced some builders to take a wait-and-see attitude (Morton et al., 2011). Exploratory work is under way to consider implementation of building codes that would focus on historic weather experience and also introduce expected future weather risks (Auld et al., 2010; Ontario Ministry of Environment, 2011). 26.7.6. Insurance 26.7.6.1. Observed Impacts Property insurance and reinsurance companies across North America experienced a significant increase in severe weather damage claims paid over the past three or four decades(Cutter and Emrich, 2005; Munich Re, 2011; Bresch and Spiegel, 2011). Most of the increase in insurance claims paid has been attributed to increasing exposure of people and assets in areas of risk (Pielke Jr et al., 2008; Barthel and Neumayer, 2012). A role for climate change has Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 not been excluded, but the increase to date in damage claims is largely due to growth in wealth and population (IPCC, 2012). Severe weather and climate risks have emerged over the past decade as the leading cost for property insurers across North America, resulting in significant change in industry practices. The price of insurance increased in regions where the risk of loss and damage has increased. Discounts have been introduced where investments in adaptation have reduced the risk of future weather losses (Mills, 2012). Further detailed discussion on the insurance sector and climate change can be found in section 10.7. 26.7.6.2. Projected Impacts Without adaptation, there is an expectation that severe weather insurance damage claims would increase significantly over the next several decades across North America (World Bank, 2010). The risk of damage is expected to rise due to continuing growth in wealth, the population living at risk, and climate change. There is also an expectation that some weather perils in North America will increase in severity, including Atlantic hurricanes and the area burned by wildfire (Karl et al., 2008; Balshi et al., 2009), and other perils in frequency, including intense rainfall events (IPCC, 2012). 26.7.6.3. Adaptation The insurance industry is one of the most studied sectors in North America in terms of climate impacts and adaptation. Most adaptation in the insurance industry has been in response to an increase in severe weather damage, with little evidence of proactive adaptation in anticipation of future climate change (Mills and Lecomte, 2006; Mills, 2007; Mills, 2009; Kunreuther and Michel-Kerjan, 2009; Leurig, 2011; Autorite des Marches Financiers, 2011; Gallagher, 2012). In addition to pricing decisions based on an actuarial analysis of historic loss experience, many insurance companies in the United States and Canada now use climate model information to help determine the prices they charge and discounts they offer. Most insurance companies have established specialized claims handling procedures for responding to catastrophic events (Kovacs, 2005; Mills, 2009). A recent study of more than 2,000 major catastrophes since 1960 found that insurance is a critical adaptive tool available to help society minimize the adverse economic consequences of natural disasters (von Peter et al., 2012). Government insurance programs for coverage of flood in the United States have been affected by recent hurricanes and previously subsidized premiums have been changed to more accurately reflect risk (Federal Emergency Management Agency, 2013). In the United States and Canada, homeowners make extensive use of insurance to manage a broad range of risks, and those with insurance recover quickly following most extreme weather events. However the majority of public infrastructure is not insured and it frequently takes more than a decade before government services fully recover. In contrast, Mexico has a well-developed program for financing the rebuilding of public infrastructure following a disaster (FONDEN) but insurance markets are only beginning to emerge for homeowners and businesses. In 2012, per capita spending on property and casualty insurance was US$2,239.20 in the United States, US$2,040.40 in Canada, and US$113.00 in Mexico (Seiler et al., 2013). Insurance companies are also working to influence the behavior of their policyholders to reduce the risk of damage from climate extremes (Kovacs, 2005; Anderson et al., 2006; Mills, 2009). For example, the industry supports the work of the Insurance Institute for Business and Home Safety in the United States, and the Institute for Catastrophic Loss Reduction in Canada, in working to champion change in the building code and communicate to property owners, governments and other stakeholders best practices for reducing the risk of damage from hurricanes, tornadoes, winter storms, wildfire, flood and other extremes. Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.8. Urban and Rural Settlements Recently a growing body of literature and national assessments have focused on climate-related impacts, vulnerabilities and risks in North American settlements (e.g., US-NCA chapters 11 and 14 and AR5 chapters 8 and 9). 26.8.1. Observed Weather and Climate Impacts Observed impacts on lives, livelihoods, economic activities, infrastructure and access to services in North American human settlements have been attributed to sea level rise (26.2.2.1), changes in temperature and precipitation, and occurrences of extreme events like heat waves, droughts and storms (Figure 26-2). Only a handful of these impacts have been attributed to anthropogenic climate change, such as shifts in Pacific Northwest marine ecosystems, which have restricted fisheries and thus affected fishing communities (U.S. Global Change Research Program, 2009). As well, (MacKendrick and Parkins, 2005; Parkins and MacKendrick, 2007; Parkins, 2008; Holmes, 2010) identified 30 communities and 25,000 families in British Columbia negatively affected by the mountain pine beetle outbreak (See 26.4.1.1). While droughts are among the more notable extreme events affecting North American urban and rural settlements recently, with severe occurrences in the Canadian Prairies causing economic and employment losses (2001-2), changes in drought frequency in North America have not been attributed to anthropogenic climate change (Figure 26-1). The 2010-2012 drought across much of the US and Northern Mexico was considered the most severe in a century (MacDonald, 2010). It affected 80% of agricultural land in the US, with 2,000 counties designated disaster zones by September (U.S. Department of Agriculture, Economic Research Service, 2012). Impacts include the loss of 3.2 million tons of maize in Mexico, placing 2.5 million at risk of food insecurity (Direccion General de Comunicacion Social, 2012). Among the most severely affected were indigenous peoples, such as the Rarámuri of Chihuahua (ibid.). Closely associated with droughts, the impacts of recent wildfires have been significant (See Box 26.2), and have intensified inequalities in vulnerability between amenity migrants and low-income residents in peri- urban areas of California and Colorado (Collins and Bolin, 2009). Other extreme-events include heat-waves, resulting in excess urban mortality (O'Neill and Ebi, 2009; Romero- Lankao et al., 2012b); and affecting infrastructure and built environments e.g., road pavement in Chicago buckled under temperatures over 100oF (CBS Chicago, 2012); in Colorado two wildfires burned over 600 homes (National Climate Data Center, 2013). Extreme storms and extreme precipitation have also impacted several North American regions (Figures 26-1 and 26- 2). Flood frequency has increased in some cities, a trend sometimes associated with more intense precipitation (e.g., Mexico City and Charlotte NC, US) (Villarini et al., 2009; Magana, 2010), while in others this trend is associated with a transition from flood events dominated by snowmelt to those caused by warm-season thunderstorms (e.g., Québec, Canada and Milwaukee, US) (Ouellet et al., 2012; Yang et al., 2013). As illustrated by Sandy (Neria and Shultz, 2012; Powell et al., 2012), storms impact human health and healthcare access (section 26.6.1.1), and impacts on infrastructure and the built environment have been costly. Heavy precipitation, storm surges, flash-floods and wind, including flooding on the US East Coast, and Midwest (2011), hurricanes and floods in the city of Villa Hermosa (Comisión Económica para América Latina y el Caribe, ) and other urban areas in southern Mexico (2004- 5), have compromised homes and businesses (Comfort, 2006; Kirshen et al., 2008; Jonkman et al., 2009; Romero- Lankao, 2010). Hurricane Wilma alone caused $1.8 billion in damage, among the biggest insurance losses in Latin American history (Galindo et al., 2009). The impacts of interacting hazards compound vulnerabilities (26.8.2). Coastal settlements are at risk from the combined occurrence of coastal erosion, health effects, infrastructure and economic damage from storm surges. Earlier thaw (Friesinger and Bernatchez, 2010), SLR, and coastal flooding have been detected along the mid- Atlantic, Gulf of Mexico, and St. Lawrence (Kirshen et al., 2008; Zavala-Hidalgo et al., 2010; Friesinger and Bernatchez, 2010; Rosenzweig et al., 2011; Tebaldi et al., 2012). Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climate impacts on the ecosystem-function and -services (e.g., water supplies, biodiversity or flood protection) provided to human settlements are another concern. While acknowledged in some places (e.g., Mexico City Climate Action Plan), they have received relatively less scholarship attention (Hunt and Watkiss, 2011). 26.8.2. Observed Factors and Processes Associated with Vulnerability Differences in the severity of climate impacts on human settlements are strongly influenced by context-specific vulnerability factors and processes (Table 26.1, Cutter et al., 2013), some of which are common to many settlements, while others are more pertinent to some types of settlements than others. Human settlements simultaneously face a multilevel array of non-climate-related hazards (e.g., economic, industrial, technological) that contribute to climate change vulnerability (McGranahan et al., 2007; Satterthwaite et al., 2007; Romero-Lankao and Dodman, 2011). In the following we highlight key sources of vulnerability for urban and rural systems. 26.8.2.1. Urban Settlements Hazard risks in urban settlements are enhanced by the concentration of populations, economic activities, cultural amenities and built environments particularly when they are in highly-exposed locations such as coastal and arid areas. Cities of concern include those in the Canadian prairies and US-Mexico border region; and major urban areas including Boston, New York, Chicago, Washington DC, Los Angeles, Villa Hermosa, Mexico City and Hermosillo (Comisión Económica para América Latina y el Caribe; Bin et al., 2007; Collins, 2008; Kirshen et al., 2008; Collins and Bolin, 2009; Gallivan et al., 2009; Rosenzweig et al., 2010; Hayhoe et al., 2010; Romero-Lankao, 2010; Wittrock et al., 2011). Risks may also be heightened by multiple interacting hazards. Slow-onset events such as urban heat-islands, for instance, interact with poor air-quality in large North American cities to exacerbate climate impacts on human health (Romero-Lankao et al., 2013a). As illustrated by recent weather events (Figure 26.2), however, hazard interactions can also follow individual, high-magnitude extreme events of short duration, with cascading effects across interconnected energy, transportation, water and health infrastructures and services to contribute to and compound urban vulnerability (Gasper et al., 2011). Wildfire vulnerability in the southwest has been compounded by peri- urban growth (Collins and Bolin, 2009; Brenkert-Smith, 2010). Under current financial constraints in many cities, climate-related economic losses can reduce resources available to address social issues, thus threatening institutional capacity and urban livelihoods (Kundzewicz et al., 2008). The urbanization process and urban built-environments of North America can amplify climate impacts as they change land-use and land-surface physical characteristics (e.g., surface albedo, Chen et al., 2011a). A 34% increase in US urban land development (Alig et al., 2004) between 1982 and 1997 had implications for water supplies and extreme event impacts. Effects on water are of special concern (section 26.3), as urbanization can enhance or reduce precipitation, depending on climate regime, geographical location and regional patterns of land, energy and water use (Cuo et al., 2009). Urbanization also has significant impacts on flood climatology through atmospheric processes tied to the Urban Heat Island (UHI), the Urban Canopy Layer (UCL), and the aerosol composition of airsheds (Ntelekos et al., 2010). The UHI can also increase health risks differentially, due to socio-spatial inequalities across and within North American cities (Harlan et al., 2008; Miao et al., 2011). Urbanization imposes path dependencies that can amplify or attenuate vulnerability (Romero-Lankao and Qin, 2011). For example, the overexploitation of Mexico City s aquifer by 19.1 - 22.2 m3/s has reduced groundwater levels and caused subsidence, undermining building foundations and infrastructure and increasing residents vulnerability to earthquakes and heavy rains (Romero-Lankao, 2010). Elements of the built-environment such as housing stock, urban form, the condition of water and power infrastructures, and changes in urban and ecological services also affect vulnerability. Large, impermeable surfaces and buildings disrupt drainage channels and accelerate run-off (Walsh et al., 2005). Damage from floods can be much more catastrophic if drainage or waste collection systems are inadequate to accommodate peak flows Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (Richardson, 2010; Sosa-Rodriguez, 2010). While many Canadian and US cities are in need of infrastructure adaptation upgrades (Doyle et al., 2008; Conrad, 2010), Mexican cites are faced with existing infrastructure deficits (Niven et al., 2010; Hardoy and Romero Lankao, 2011), and high levels of socio-spatial segregation (Smolka and Larangeira, 2008) (section 26.7). Recent weather hazards (Figure 26-2) illustrate that economic activities and highly-valued physical capital of cities (real estate, interconnected infrastructure systems) are very sensitive to climate-related disruptions that can result in high impacts; activities in some urban areas are particularly exposed to key resource constraints (e.g., water in the US-Mexico Border; oil industry in Canada, US and Mexico; Levy et al., 2010; Conrad, 2010); others are dependent upon climate-sensitive sectors (e.g., tourism) (Lal et al., 2011). Disruptions to production, services and livelihoods, and changes in the costs of raw materials also impact the economic performance of cities (Hunt and Watkiss, 2011). Cities are relatively better endowed than rural populations with individual and neighborhood assets such as income, education, quality of housing and access to infrastructure and services that offer protection from climate hazards. However, intra-urban socio-spatial differences in access to these assets shape response capacities (Harlan and Ruddell, 2011; Romero-Lankao et al., 2013a). All this means that class and socio-spatial segregation are key determinants not only of vulnerability but also of inequalities in risk generation and distribution within cities. Economic elites are better positioned to access the best land and enjoy the rewards of environmental amenities such as clean air, safe drinking water, open space, and tree shade (Morello-Frosch et al., 2002; Harlan et al., 2006; Harlan et al., 2008; Ruddell et al., 2011). Although wealthy sectors are moving into risk prone coastal and forested areas (Collins, 2008), and certain hazards (air pollution) affect both rich and poor alike (Romero-Lankao et al., 2013a), climate risks tend to be disproportionally borne by the poor or otherwise marginalized populations (Cutter et al., 2008; Collins and Bolin, 2009; Romero-Lankao, 2010; Wittrock et al., 2011). In some cities, marginalized populations are moving to peri-urban areas with inadequate services, a portfolio of precarious livelihood mechanisms, and inappropriate risk-management institutions (Collins and Bolin, 2009; Eakin et al., 2010; Monkkonen, 2011; Romero-Lankao et al., 2012a). Although cities have comparatively higher access than rural municipalities to determinants of institutional capacity such as human resources and revenue pools, their governance arrangements are often hampered by jurisdictional conflicts, asymmetries in information and communication access, fiscal constraints on public services including emergency personnel, and top-down decision making. These governance issues exacerbate urban vulnerabilities and constrain urban adaptation planning (Carmin et al., 2012; Romero-Lankao et al., 2013a). 26.8.2.2. Rural Settlements The legacy of previous and current stresses contributing to rapid population growth or loss in North American rural communities, reduced employment, or degradation of local knowledge systems, can increase vulnerability (Brklacich et al., 2008; Coles and Scott, 2009; McLeman, 2010). North American rural communities have a higher proportion of lower income and unemployed populations and higher poverty than cities (Whitener and Parker, 2007; Lal et al., 2011; Skoufias et al., 2011). 55% of Mexico s rural residents live in poverty, and the livelihood of 72% of these is in farming (Saldana-Zorrilla, 2008). US and Canadian rural communities have older populations (McLeman, 2010) and lower education levels (Lal et al., 2011). Indigenous communities have lower education levels, and high levels of poverty, but are younger than average populations (Downing and Cuerrier, 2011). The legacy of their colonial history, furthermore, has stripped Indigenous communities of land and many sources of social and human capital (Brklacich et al., 2008; Hardess et al., 2011). Conversely, rural and indigenous community members possess valuable local and experiential knowledge regarding regional ecosystem services (Nakashima et al., 2011). Rural economies have limited economic diversity and relatively high dependence on climate-sensitive sectors (Johnston et al., 2008; Natural Resources Canada, 2008; Molnar, 2010); they are sensitive to climate-induced reductions in resource supply and productivity, in addition to direct exposure to climate hazards (Daw et al., 2009). Single-sector economic dependence contributes significantly to vulnerability (Cutter et al., 2003). Engagement in export markets presents opportunity but also exposure to economic volatility (Eakin, 2006; Saldana-Zorilla and Sandberg, 2009), and economic downturns take attention away from climate change adaptation. Farming and fishing Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 provide both economic and food security, the impacts of climate thus posing a double threat to livelihood (Badjek et al., 2010), particularly among women (Bee et al., 2013). Inter-related factors affecting vulnerability in forestry and fishing communities include over-harvesting, and the cumulative environmental effects of multiple land use activities (Brklacich et al., 2008). Many tourism-based communities are dominated by seasonal economies and low- wage, service-based employment (Tufts, 2010), and small businesses that lack resources for emergency planning (Hystad and Keller, 2006; Hystad and Keller, 2008). Non-renewable resource industries are sensitive to power, water, and transportation disruptions associated with hazards. Geographic isolation can be a key source of vulnerability for rural communities in North America, imposing long commutes to essential services like hospitals, and non-redundant transportation corridors that can be compromised during extreme events (Chouinard et al., 2008). Many Indigenous communities are isolated, raising the costs and limiting the diversity of imported food, fuel and other supplies, rendering the ability to engage in subsistence harvesting especially critical for both cultural and livelihood wellbeing (Andrachuk and Pearce, 2010; Hardess et al., 2011). Many indigenous peoples also maintain strong cultural attachment to ancestral lands, and thus are especially sensitive to declines in the ability of that land to sustain their livelihoods and cultural wellbeing (Downing and Cuerrier, 2011). Rural physical infrastructure is often inadequate to meet service needs or is in poor condition (McLeman and Gilbert, 2008; Krishnamurthy et al., 2011), especially for Indigenous communities (section 26.9, Brklacich et al., 2008; Hardess et al., 2011; Lal et al., 2011). A lack of redundant power and communication services can compromise hazard response capacity. 26.8.3. Projected Climate Risks on Urban and Rural Settlements Urbanization, migration, economic disparity, and institutional capacity will influence future impacts and adaptation to climate change in North American human settlements (section 26.2.1). Water related concerns are assessed in 26.3.2.1 and 26.3.2.3). We describe below a variety of future climate risks identified in the literature, many of which focus on cities (Chapters 8 and 9) and, with the exception of larger centers such as New York and Boston, are qualitative in nature (Hunt and Watkiss, 2011). This is due in part to the difficulty in downscaling the prediction of trends in key trends in climate parameters to an appropriate scale Model-based sea-level-rise (SLR) projections of future risks to cities are characterized by large uncertainties due to global factors (e.g., the dynamics of polar ice-sheets WGI) and regional factors (e.g., regional shifts in ocean circulation, high of the adjacent ocean and local land high) (Blake et al., 2011) (WGI chapter 3). The latter will determine differential SLR impacts on regional land development of coastal settlements (U.S. Government Accountability Office, 2007; Yin et al., 2009; Sobel et al., 2010; Conrad, 2010; Millerd, 2011), making some areas particularly vulnerable to inundation (Cooper and Sehlke, 2012). SLR can also exacerbate vulnerability to extreme events such as hurricanes (Frazier et al., 2010). Temperature increases would lead to additional health hazards. Baseline warmer temperatures in cities are expected to be further elevated by extreme heat events whose intensity and frequency is projected to increase during the 21st century (section 26.2.2), particularly in northern mid-latitude cities (Jacob and Winner, 2009). Participation in some outdoor activities would also increase as a result of projected increases in warm days (Scott and McBoyle, 2007). Projected snowfall declines in Canada and the Northeast US would reduce length of winter sport seasons and thus affect the economic wellbeing of some communities (McBoyle et al., 2007; Scott et al., 2008b). Any increase in frequency of extreme events, such as intense precipitation, flooding and prolonged dry periods would affect particularly the populations, economic activities, infrastructures and services on coasts, flood-prone deltas and arid regions (Nicholls et al., 2008; Kirshen et al., 2008; Richardson, 2010; Weiss et al., 2011). For example, by the end of this century, New York City is projected to experience nearly twice as many extreme precipitation days compared to today (A2, mean ensemble of 17 models). (Ntelekos et al., 2010; Cayan et al., 2010) project an increase in the number and duration of droughts in the southwest US, with most droughts expected to last over five years by 2050 (GDFL CM2.1 and CNRM CM3, A2 and B1). Assuming no adaptation, total loses from Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 river flooding in metropolitan Boston are estimated to exceed $57 billion by 2100, of which $26 billion is attributed to climate change (Nicholls et al., 2008; Kirshen et al., 2008; Richardson, 2010; Weiss et al., 2011). Future climate risks on lives and livelihoods have been relatively less studied. A handful of studies focused on forestry are notable, indicating potentially substantial shifts in livelihood options without adaptation. Sohngen and Sedjo (Sohngen and Sedjo, 2005) estimate losses from climate change in the Canadian/US timber sector of $1.4 $2.1 billion per year over the next century. Anticipated future supply reductions in British Columbia as a consequence of the pine beetle outbreak vary from 10 to 62% (Patriquin et al., 2007). Substantial declines in suitable habitat for valued tree species in Mexico have been projected (Gómez-Mendoza and Arriaga, 2007; Gomez-Diaz et al., 2011). Scholars are starting to project future risks from interacting hazards. For instance, by 2070 with a 0.5m rise in sea- level and under scenarios of socioeconomic growth, storm surges and subsidence, populations at risk in New York, Miami and New Orleans might increase three-fold, while asset exposure will increase more than 10-fold (Hanson et al., 2011). Essential infrastructure and services are key concerns (sections 26.3 and 26.7). Increased occurrence of drought affecting water availability is projected for southwestern US/Northern Mexico, the southern Canadian Prairies and central Mexico, combined with projected increases in water demand due to rapid population growth and agriculture (Schindler and Donahue, 2006; MacDonald, 2010; Lal et al., 2011). Using A1B and A2 scenarios, (Escolero- Fuentes et al., 2009) projected that by 2050, Mexico City and its watersheds will experience a more intense hydrological cycle and a reduction of between 10 to 17% in per capita available water. Sea-level rise is predicted to threaten water and electricity infrastructure with inundation and increasing salinity (Sharp, 2010). 26.8.4. Adaptation 26.8.4.1. Evidence of Adaptation 26.8.4.1.1. What are populations doing? Autonomous Adaptation As illustrated by recent extreme events (Figure 26-2), individuals and households in North America have not only be affected by extremes, but have also been responding to climate impacts mostly through incremental actions, for example by purchasing additional insurance, or reinforcing homes to withstand extreme weather (Simmons and Sutter, 2007; Romero-Lankao et al., 2012a). Some individuals respond by diversifying livelihoods (Newland et al., 2008; Rose and Shaw, 2008) or migrating (See 26.1.1) (Black et al., 2011). The propensity to respond to climate and weather hazards is strongly influenced not only by access to household assets, but also by community and governmental support. The emergency response to Sandy illustrates this. Although New York and New Jersey witnessed vivid scenes of medical humanitarianism, because of inadequate communication and coordination among agencies, public health support did not always reach those most in need (Abramson and Redlener, 2013). The perceived risks of climate change among individuals are equally important. Strong attachment to place and occupation may motivate willingness to support incremental adaptation, enhance coping capacity and foster adaptive learning (Collins and Bolin, 2009; Romero-Lankao, 2010; Aguilar and Santos, 2011; Wittrock et al., 2011). They have also been found to serve as barriers to transformational adaptation (Marshall et al., 2012). Residents of the US stand out in international research as holding lower levels of perceived risk of climate change (AXA Ipsos, 2012), which may limit involvement in household-level adaptation, or support for public investments in adaptation. Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.8.4.1.2. What are governments doing? Planned Adaptation Leadership in adaptation is far more evident locally than at other tiers of government in North America (Richardson, 2010; Vasseur, 2011; Vrolijks et al., 2011; Henstra, 2012; Carmin et al., 2012). Few municipalities have moved into the implementation stage, however; most programs are in the process of problem diagnosis and planning (Perkins et al., 2007; Moser and Satterthwaite, 2008; Romero-Lankao and Dodman, 2011). Systematic assessments of vulnerability are rare, particularly in relation to population groups (Vrolijks et al., 2011). Surveys of municipal leaders showed adaptation is rarely incorporated into planning, due to lack of resources, information and expertise (Horton and Richardson, 2011), and the prevalence of other issues considered higher priority, suggesting the need for subnational and federal-level facilitation in the form of resources and enabling regulations . Climate change policies have been motivated by concerns for local economic or energy security and the desire to play leadership roles (Rosenzweig et al., 2010; Anguelovski and Carmin, 2011; Romero-Lankao et al., 2013a). Some policies constitute integrated strategies (New York) (Perkins et al., 2007; Rosenzweig et al., 2010), and coordinated participation of multiple municipalities (Vancouver) (Richardson, 2010). Sector-specific climate risk management plans have also emerged (e.g., water conservation in Phoenix, US and Regina, Canada; wildfire protection in Kamloops, Canada and Boulder, US). Municipalities affected by the mountain pine beetle have taken many steps toward adaptation (Parkins, 2008), and coastal communities in eastern Canada are investing in saltwater marsh restoration to adapt to rising sea levels (Marlin et al., 2007). Green roofs, forest thinning and urban agriculture have all been expanding (Chicago, New York, Kamloops, Mexico City), as have flood protection (New Orleans, Chicago), private and governmental insurance policies (section 26.10, Browne and Hoyt, 2000; Ntelekos et al., 2010), safe saving schemes (common in Mexico), air pollution controls (Mexico City), and hazard warning systems (Collins and Bolin, 2009; Coffee et al., 2010; Romero-Lankao, 2010; Aguilar and Santos, 2011). 26.8.4.2. Opportunities and Constraints Adaptation in human settlements is influenced by local access to resources, political will and the capacity for institutional-level attention and multilevel/multisectoral coordination (Burch, 2010; Romero-Lankao et al., 2013a) (discussed further in the next section). 26.8.4.2.1. Adaptation is path-dependent Adaptation options are constrained by past settlement patterns and decisions. The evolution of cities as economic hubs, for example, affects vulnerability and resilience (Leichenko, 2011). Urban expansion into mountain, agricultural, protected and otherwise risk-prone areas (Boruff et al., 2005; McGranahan et al., 2007; Collins and Bolin, 2009; Conrad, 2010) invariably alters regional environments. Development histories foreclose some resilience pathways. Previous water development, for example, can result in irreversible overexploitation and degradation of water resources. 26.8.4.2.2. Institutional capacity At all levels of governance, adaptation in North America is affected by numerous determinants of institutional capacity. Three have emerged in the literature as particularly significant challenges for urban and rural settlements: Economic Resources: Rural communities face limited revenues combined with higher costs of supplying services (Williamson et al., 2008; Posey, 2009). Small municipal revenue pools translate into fiscal constraints necessary to support public services, including emergency personnel and health care (Lal et al., 2011). Although large cities tend to have greater fiscal capacity, most do not receive financial support for adaptation (Carmin et al., 2012), yet face the risk of higher economic losses. Information and social capital: Differences in access and use of information, and capacity for learning and innovation, affect adaptive capacity (Romero-Lankao et al., 2013a). Levels of knowledge and prioritization can be low among municipal planners. Information access can be limited, even among environmental Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 planners (Picketts et al., 2012). The relationship between trust and participation in support networks (social capital) and adaptive capacity is generally positive, however strong social bonds may support narratives that under-estimate climate risk (Wolf et al., 2010; Romero-Lankao et al., 2012b). Participation: Considering the overlap among impacts and sources of vulnerability in North American human settlements, long-term effectiveness of local adaptation hinges upon inclusion of all stakeholders. Stakeholder involvement lengthens planning time frames, may elicit conflicts, and power relationships can constrain access (Few et al., 2007; Colten et al., 2008). However, effective stakeholder engagement has tremendously enhanced adaptation planning, eliciting key sources of information regarding social values, securing legitimacy (Aguilar and Santos, 2011), and fostering adaptive capacity of involved stakeholders. _____ START BOX 26-3 HERE _____ Box 26-3. Climate Responses in Three North American Cities With populations of 20.5, 14 and 2.3 million people, respectively, the metropolitan areas of Mexico City, New York, and Vancouver are facing multiple risks that climate change is projected to aggravate. These risks range from sea level rise, coastal flooding and storm surges in New York and Vancouver to heat waves, heavy rains and associated flooding, air pollution, and heat-island effects in all the three cities (Rosenzweig and Solecki, 2010; Leon and Neri, 2010; City of Vancouver, 2012). Many of these risks result not only from long-term global and regional processes of environmental change, but also from local changes in land and water uses and in atmospheric emissions induced by urbanization (Romero-Lankao, 2010; Leon and Neri, 2010; Kinney et al., 2011; Solecki, 2012). The three cities have been frontrunners in the climate arena. In Mexico City, the Program of Climate Action 2008 2012 (PAC) and the 2011 Law for Mitigation and Adaptation to Climate Change are parts of a larger 15-year Green Agenda, with most of designated funds committed to reducing 7 million tonnes of CO2 equivalent by 2012 (Romero-Lankao et al., 2013). New York City s and Vancouver s Plans are similarly mitigation-centred. As of 2007 New-York s long-term sustainability plan included adaptation (Solecki, 2012; Ray et al., 2013), while Vancouver launched its municipal adaptation plan in July 2012. The shifts in focus from mitigation to adaptation have followed as it has become increasingly clear that even if mitigation efforts are wholly successful, some adverse impacts due to climate change are unavoidable. Urban leaders in all three cities have emerged as global leaders in sustainability. Mayor Bloomberg of New York; Mayor Ebrard of Mexico City, and David Cadman of Vancouver have, respectively, led the C40, World Mayors Council on Climate Change and International Council for Local Environmental Initiatives (ICLEI). Scientists, private sector actors and nongovernmental organizations have been of no lesser importance. To take advantage of a broad-based interaction between various climate change actors, Mexico City has set up a Virtual Climate Change Centre to serve as a repository of knowledge, models and data on climate change impacts, vulnerability and risks (Romero-Lankao et al., 2013). Information sharing by climate change actors has also taken place in New York, where scientists, and insurance and risk management experts have served on the Panel on Climate Change to advise the city on the science of climate change impacts and protection levels specific to the city s critical infrastructure (Solecki, 2012): 564). The climate plans of the three cities are far reaching, including mitigation and adaptation strategies related to their sustainability goals. The three cities emphasize different priorities in their climate action plans. Mexico City seeks to reduce water and transportation emissions through such actions as improvements in infrastructure and changes in the share of public transport. Vancouver has prioritized the separation of sanitary and stormwater systems, yet this adaptation is not expected to be complete until 2050 (City of Vancouver, 2012). It will also take New York much time, money and energy to expand adaptation strategies beyond the protection of water systems to include all essential city infrastructure (Ray et al., 2013). Overall, few proposed actions will result in immediate effects, and instead call for additional planning, highlighting the significant effort necessary for comprehensive responses. Overall, adaptation planning in the three cities faces many challenges. In all three regions, multi-jurisdictional governance structures with differing approaches to climate change challenge the ability for coordinated responses (Solecki, 2012; Romero-Lankao et al., 2013). Conflicts in priorities and objectives between various actors and Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 sectors are also prevalent (Burch, 2010). For instance, authorities in Mexico City concerned with avoiding growth into risk-prone and conservation areas (Aguilar and Santos, 2011) compete for regulatory space within a policy agenda that is already coping with a wide range of economic and developmental imperatives (Romero-Lankao et al., 2013). Climate responses require new types of localized scientific information, such as vulnerability analyses and flood risk assessments, which are not always available (Romero-Lankao et al., 2012a; Ray et al., 2013). Little is known, for instance, about how to predict and respond to common and differential levels of risk experienced by different human settlements. Comprehensive planning is still limited as well. For example, although scholarship exists on disparities in household- and population-level vulnerability and adaptive capacity (Villeneuve and Burnett, 2003; Cutter et al., 2003; Douglas et al., 2012; Romero-Lankao et al., 2013b), equity concerns have received relatively less attention by either of the three cities. Even when local needs are identified, such as the need to protect higher risk homeless and low-income populations (Vancouver), they are often not addressed in action plans. _____ END BOX 26-3 HERE _____ 26.9. Federal and Subnational Level Adaptation Along with many local governments (section 26.8.4), federal, and subnational tiers of government across North America are developing climate change adaptation plans. These initiatives, which began at the subnational levels (e.g., (Nunavut Department of Sustainable Development, 2003), appear to be preliminary and relatively little has been done to implement specific measures. 26.9.1. Federal Level Adaptation All three national governments are addressing adaptation to some extent, with a national strategy and a policy framework (Mexico), a federal policy framework (Canada), and the United States having delegated all federal agencies to develop adaptation plans. In 2005, the Mexican government created the Inter-Secretarial Commission to Climate Change (CICC Comisión Inter-Secretarial de Cambio Climático) to coordinate national public policy on climate change (Comisión Inter- Secretarial de Cambio Climático, 2005; Sosa-Rodriguez, 2013).The government s initiatives are being delivered through the National Strategy for Climate Change 2007-2012 (Intersecretarial Commission on Climate Change, 2007) and, the Special Programme on Climate Change 2009-2012, which identify priorities in research, cross- sectoral action such as developing early warning systems, and capacity development to support mitigation and adaptation actions (Comisión Inter-Secretarial de Cambio Climático, 2009). The Policy Framework for Medium Term Adaptation (Consejo Intersecretarial de Cambio Climático, 2010) aims at framing a single national public policy approach on adaptation with a time-horizon up to 2030. The General Law of Climate Change requires state governments to implement mitigation and adaptation actions (Diario Oficial de la Federación, 2012). Canada is creating a Federal Adaptation Policy Framework intended to mainstream climate risks and impacts into programs and activities to help frame government priorities (Environment Canada, 2011a). In 2007, the federal Government made a four-year adaptation commitment to develop six Regional Adaptation Collaboratives (RAC) in provinces across Canada, ranging in size and scope, from flood protection and drought planning, to extreme weather risk management; and assessing the vulnerability of Nunavut s mining sector to climate change (Natural Resources Canada, 2011). In 2011, the federal government renewed financial support for several adaptation programs and provided new funding to create a Climate Adaptation and Resilience Program for Aboriginals and Northerners, and Enhancing Competitiveness in a Changing Climate program (Environment Canada, 2011b). Canada recently launched an Adaptation Platform to advance adaptation priorities across the country (Natural Resources Canada, 2013). Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The US government embarked in 2009 on a government-wide effort to have all federal agencies address adaptation; to apply understanding of climate change to agency missions and operations; to develop, prioritize, and implement actions; and to evaluate adaptations and learn from experience. (The White House, 2009; Bierbaum et al., 2012). A 2013 plan issued by the President enhanced the US government effort supporting adaptation (Executive Office of the President, 2013). The US Government provides technical and information support for adaptation by non-federal actors, but does not provide direct financial support for adaptation (Parris et al., 2010). Some federal agencies took steps to address climate change adaptation prior to this broader interagency effort. In 2010, the US Department of Interior created Climate Science Centers to integrate climate change information and management strategies in eight regions and 21 Landscape Conservation Cooperatives (Secretary of the Interior, 2010), while the US Environmental Protection Agency s Office of Water developed a climate change strategy (U.S. Environmental Protection Agency, National Water Program, 2011). 26.9.2. Subnational Level Adaptation A number of states and provinces in all three countries have developed adaptation plans. For example, in Canada, Quebec's 2013-2020 adaptation strategy outlines 17 objections covering a number of managed sectors and ecosystems (Government of Quebec, 2012). British Columbia is modernizing its Water Act to alter water allocation during drought to reduce agricultural crop and livestock loss and community conflict, while protecting aquatic ecosystems (British Columbia Ministry of the Environment, 2010). In the US California was the first state to publish an adaptation plan calling for a 20% reduction in per capita water use by 2020 (California Natural Resources Agency, 2009). Maryland first developed a plan on coastal resources and then broadened it to cover human health, agriculture, ecosystems, water resources, and infrastructure (Maryland Commission on Climate Change Adaptation and Response Working Group, 2008; Maryland Department of the Environment on behalf of the Maryland Commission on Climate Change, 2010). The State of Washington is addressing environment, infrastructure, and communities; human health and security; ecosystems, species, and habitat; and natural resources (Washington State Built Environment: Infrastructure & Communities Topic Advisory Group, 2011; Washington State Human Health and Security Topic Advisory Group, 2011; Washington State Species, Habitats and Ecosystems Topic Advisory Group, 2011; Washington State Natural Resources Working Lands and Waters Topic Advisory Group, 2011). Of the three national governments, only Mexico requires that states develop adaptation plans. In Mexico, seven of 31 states, Veracruz, Mexico City, Nuevo León, Guanajuato, Puebla, Tabasco, and Chiapas, have all developed their State Programmes for Climate Change Action (Programas Estatales de Acción ante el Cambio Climático - PEACC), while Baja California Sur, Hidalgo, and Campeche are in the final stage and 17 states are still in the planning and developing stage (Instituto de Ecología del Estado de Guanajuato, 2011). The proposed adaptation actions focus mainly on: 1) reducing physical and social vulnerability of key sectors and populations; 2) conservation and sustainable management of ecosystems, biodiversity, and ecosystem services; 3) developing risk management strategies; 4) strengthening water management; 5) protecting human health, and; 6) improving current urban development strategies, focusing on settlements and services, transport and land use planning. 26.9.3. Barriers to Adaptation Chapter 16 provides a more in-depth discussion on adaptation barriers and limits. Adaptation plans tend to exist as distinct documents and are often not integrated into other planning activities (Preston et al., 2011). Most adaptation activities have only involved planning for climate change rather than specific actions, and few measures have been implemented (Preston et al., 2011; Bierbaum et al., 2012). Even though Canada and the US are relatively well-endowed in their capacity to adapt, there are significant constraints on adaptation, with financing being a significant constraint in all three countries (Carmin et al., 2012). Barriers include legal constraints (e.g., (Jantarasami et al., 2010) lack of coordination across different jurisdictions Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (Smith et al., 2009; National Research Council, 2010; Instituto Nacional de Ecología y Cambio Climático, 2012b), leadership (Smith et al., 2009; Moser and Ekstrom, 2010), and divergent perceptions about climate change (Bierbaum et al., 2012; Moser, 2013). Although obtaining accurate scientific data was ranked less important by municipalities(Carmin et al., 2012), an important constraint is lack of access to scientific information and capacity to manage and use it (Moser and Ekstrom, 2010; Instituto Nacional de Ecología y Cambio Climático, 2012b). Adaptation activities in developed countries such as the US tend to address hazards and propose adaptations that tend to protect current activities rather than facilitate long term change. In addition, the adaptation plans generally do not attempt to increase adaptive capacity (Eakin and Patt, 2011). However, making changes to institutions needed to enable or promote adaptations can be costly (Marshall, 2013). Although multilevel and multisectoral coordination is a key component of effective adaptation, it is constrained by factors such as mismatch between climate and development goals, political rivalry, and lack of national support to regional and local efforts (Brklacich et al., 2008; Sander-Regier et al., 2009; Brown, 2009; Sydneysmith et al., 2010), (Craft and Howlett, 2013; Romero-Lankao et al., 2013a). Traditionally, environmental or engineering agencies are responsible for climate issues (e.g., Mexico City, Edmonton and London, Canada), but have neither the decision making power nor the resources to address all dimensions involved. Adaptation planning requires long- term investments by government, business, grassroots organizations and individuals (e.g., Romero-Lankao, 2007; Croci et al., 2010; Sarah, 2010; Richardson, 2010). 26.9.4. Maladaptation, Trade-Offs, and Co-Benefits Adaptation strategies may introduce trade-offs or maladaptive effects for policy goals in mitigation, industrial development, energy security, and health (Hamin and Gurran, 2009; Laukkonen et al., 2009). Snow-making equipment, for example, mediates snowpack reductions, but has high water and energy requirements (Scott et al., 2007). Irrigation and air conditioning have immediate adaptive benefits for North American settlements, but are energy-consumptive. Sea walls protect coastal properties, yet negatively affect coastal processes and ecosystems (Richardson, 2010). Conventional sectoral approaches to risk management and adaptation planning undertaken at different temporal and spatial scales have exacerbated vulnerability in some cases e.g., peri-urban areas in Mexico (Eakin et al., 2010; Romero-Lankao, 2012). Approaches that delegate response planning to residents in the absence of effective knowledge exchange have resulted in maladaptive effects (Friesinger and Bernatchez, 2010). Other strategies offer synergies and co-benefits. Policies addressing air pollution (Harlan and Ruddell, 2011) or housing for the poor, particularly in Mexico (Colten et al., 2008), can often be adapted at low or no cost to fulfill adaptation and sustainability goals (Badjek et al., 2010). Efforts to temper declines in production or competitiveness in rural communities could involve mitigation innovations, including carbon sequestration forest plantations (Holmes, 2010). Painting roofs white reduces the effects of heat and lowers energy demand for cooling (Akbari et al., 2009). Adaptation planning can be greatly enhanced by incorporating regionally or locally-specific vulnerability information (Clark et al., 1998; Barsugli et al., 2012; Romsdahl et al., 2013). Methods for mapping vulnerability have been improved and effectively utilized (Romero-Lankao et al., 2013b). Similarly, strategies supporting cultural preservation and subsistence livelihood needs among Indigenous peoples would enhance adaptation (Ford et al., 2010b), as would integrating traditional culture with other forms of knowledge, technologies, education and economic development (Hardess et al., 2011). Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 26.10. Key Risks, Uncertainties, Knowledge Gaps, and Research Needs 26.10.1. Key Multi-Sectoral Risks We close this chapter with our assessment of key current and future regional risks from climate change with an evaluation of the potential for risk reduction through adaptation (Table 26-1). Two of the three examples, wildfires and urban floods, illustrate that multiple climate drivers can result in multiple impacts (e.g., loss of ecosystems integrity, property damage and health impacts due to wildfires and urban floods). The three risks evaluated in Table 26-1 also show that relative risks depend on the context specific articulation and dynamics of such factors as The magnitude and rate of change of relevant climatic and non-climatic drivers and hazards. For instance, the risk of urban floods depends not only on global climatic conditions (current versus future global mean temperatures of 2 C and 4 C), but also on urbanization, a regional source of hazard risk that can enhance or reduce precipitation, as it affects the hydrologic cycle and, hence, has impacts on flood climatology (section 26.8.2.1); The internal properties and dynamics of the system being stressed. For example, some ecosystems are more fire-adapted than others. Some populations are more vulnerable to heat-stress because of age, pre-existing medical conditions, working conditions and lifestyles (e.g., outdoor workers, athletes); Adaptation potentials and limits. For example, while residential air conditioning (A/C) can effectively reduce health risk, availability and usage of A/C is often limited among the most vulnerable individuals. Furthermore, A/C is sensitive to power failures and its use has mitigation implications. [INSERT TABLE 26-1 HERE Table 26-1: Key risks from climate change and the potential for risk reduction through adaptation. Key risks are identified based on assessment of the literature and expert judgments made by authors of this chapter, with supporting evaluation of evidence and agreement in the referenced chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels are presented for the near-term era of committed climate change (here, for 2030-2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the longer-term era of climate options (here, for 2080-2100), for global mean temperature increase of 2°C and 4°C above preindustrial levels. For each timeframe, risk levels are estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols.] The judgments about risk conveyed by the Table 26-1 are based on assessment of the literature and expert judgment by chapter authors living under current socio-economic conditions. Therefore, risk levels are estimated for each timeframe, assuming a continuation of current adaptation potentials and constraints. Yet over the course of the 21st century, socioeconomic and physical conditions can change considerably for many sectors, systems and places. The dynamics of wealth generation and distribution, technological innovations, institutions, even culture, can substantially affect North American levels of risk tolerance within the social and ecological systems considered in the Table (see also Box TS.8). 26.10.2. Uncertainties, Knowledge Gaps, and Research Needs The literature on climate impacts, adaptation and vulnerability in North America has grown considerably, as has the diversity of sectors and topics covered (e.g., urban and rural settlements, food security, and adaptation at local, state and national levels). However, limitations in the topical and geographical scope of this literature are still a challenge (e.g., more studies have focused on insurance than on economic sectors such as industries, construction and transportation ). It is also challenging to summarize results across many studies and identify trends in the literature when there are differences in methodology, theoretical frameworks and causation narratives (e.g., between outcome and contextual approaches) making it hard to compare apples to oranges (Romero-Lankao et al., 2012b) While the US and Canada have produced large volumes of literature, Mexico lags well behind. It was, therefore, difficult to devote equal space to observed and projected impacts, vulnerabilities and adaptations in Mexico in comparison with Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 its Northern neighbors. With its large land area, population and important, albeit understudied, climate change risks and vulnerabilities, more climate change research focusing on Mexico is direly needed. The literature on North America tends to be dominated by sector level analyses. Yet, climate change interacts with other physical and social processes to create differential risks and impact levels. These differences are mediated by context-specific physical and social factors shaping the vulnerability of exposed systems and sectors. Furthermore, while studies often focus on isolated sectorial effects, impacts happen in communities, socio-ecologic systems and regions, and shocks and dislocations in one sector or region often affect other sectors and regions due to social and physical interdependencies. This point is illustrated by our border region and wildfire boxes and the human settlements section, which discuss place-based impacts, vulnerabilities and adaptations. Unfortunately, literature using placed-based or integrated approaches to these complexities is limited. Indeed although in early drafts the authors of this chapter attempted to put more emphasis on place-based analysis and comparisons, the literature was inadequate to support such an effort. The IPCC includes chapters on continents and large regions to make it possible to assess how multiple climate change impacts can affect these large areas. However, this macro view gives insufficient detail on context specific local impacts and risks, missing the on-the-ground reality that the effects of climate change are and will be experienced at much smaller scales, and those smaller scales are often where meaningful mitigation and adaptation actions can be generated. In order to give local actors relevant information on which to base these local actions, more research is needed to better understand the local and regional effects of climate change across sectors. Frequently Asked Questions FAQ 26.1: What impact is climate having on North America? [to remain at the end of the chapter] Recent climate changes and extreme events demonstrate clear impacts of climate-related stresses in North America (high confidence). There has been increased occurrence of severe hot weather events over much of the US and increases in heavy precipitation over much of North America (high confidence). Such events as droughts in northern Mexico and south-central US, floods in Canada, and hurricanes such as Sandy, demonstrate exposure and vulnerability to extreme climate (high confidence). Many urban and rural settlements, agricultural production, water supplies, and human health have been observed to be vulnerable to these and other extreme weather events (Figure 26-2). Forest ecosystems have been stressed through wildfire activity, regional drought, high temperatures, and infestations, while aquatic ecosystems are being affected by higher temperatures and sea level rise. Many decision makers, particularly in the United States and Canada, have the financial, human and institutional capacity to invest in resilience, yet a trend of rising losses from extremes has been evident across the continent (Figure 26-2), largely due to socio-economic factors, including a growing population, equity issues and increased property value in areas of high exposure. In addition, climate change is very likely to lead to more frequent extreme heat events and daily precipitation extremes over most areas of North America, more frequent low snow years, and shifts towards earlier snowmelt runoff over much of the western US and Canada (high confidence). These changes combined with higher sea levels and associated storm surges, more intense droughts, and increased precipitation variability are projected to lead to increased stresses to water, agriculture, economic activities and urban and rural settlements (high confidence). FAQ 26.2: Can adaptation reduce the adverse impacts of climate in North America? [to remain at the end of the chapter] Adaptation including land use planning, investments in infrastructure, emergency management, health programs, and water conservation has significant capacity to reduce risks from current climate and climate change (Figure 26-3). There is increasing attention to adaptation among planners at all levels of government but particularly at the municipal level, with many jurisdictions engaging in assessment and planning processes. Yet, there are few documented examples of implementation of proactive adaptation and these are largely found in sectors with longer- term decision-making, including energy and public infrastructure (high confidence). Adaptation efforts have revealed the significant challenges and sources of resistance facing planners at both the planning and implementation stages, particularly the adequacy of informational, institutional, financial and human resources, and lack of political will (medium confidence). Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 While there is high capacity to adapt to climate change across much of North America, there are regional and sectoral disparities in economic resources, governance capacity, and access to and ability to utilize information on climate change which limit adaptive capacity in many regions and among many populations such as the poor and indigenous communities. For example, there is limited capacity for many species to adapt to climate change, even with human intervention. At lower levels of temperature rise, adaptation has high potential to off-set projected declines in yields for many crops, but this effectiveness is expected to be much lower at higher temperatures. The risk that climate stresses will cause profound impacts on ecosystems and society including the possibility of species extinction or severe adverse socio-economic shocks highlights limits to adaptation. References Abramson, D.M. and I.E. Redlener, 2013: Hurricane Sandy: lessons learned, again. Disaster Medicine and Public Health Preparedness, 6(4), 328-329. Acclimatise, 2009: Building business resilience to inevitable climate change. In: Carbon project disclosure report 2008. FTSE 350, Oxford. Aguilar, A.G. and C. Santos, 2011: Informal settlements' needs and environmental conservation in Mexico City: An unsolved challenge for land-use policy. Land use Policy, 28(4), 649-662. Aguilera, E., L. Lassaletta, A. Gattinger, and B.S. Gimeno, 2013: Managing soil carbon for climate change mitigation and adaptation in Mediterranean cropping systems: A meta-analysis. Agriculture, Ecosystems & Environment, 168, 25-36. Ahmed, S.A., N.D. Diffenbaugh, and T. Hertel, 2009: Climate volatility deepens poverty vulnerability in developing countries. Environmental Research Letters, 4(3), 034004. Akbari, H., S. Menon, and A. Rosenfeld, 2009: Global cooling: increasing world-wide urban albedos to offset CO 2. Climatic Change, 94(3), 275-286. Alam, M.M., C. Siwar, R.I. Molla, B. Talib, and M.E.B. Toriman, 2012: Paddy farmers' adaptation practices to climatic vulnerabilities in Malaysia. Mitigation and Adaptation Strategies for Global Change, 17, 415-423. Alexander et al., 2006: Global observed changes in daily climate extremes of temperature and precipitation. Journal of Geophysical Research, 111(D5), D05109. Alig, R.J., J.D. Kline, and M. Lichtenstein, 2004: Urbanization on the US landscape: looking ahead in the 21st century. Landscape and Urban Planning, 69(2), 219-234. Allen, C.D., A. Macalady, H. Chenchouni, D. Bachelet, N. McDowell, M. Vennetier, P. Gonzales, T. Hogg, A. Rigling, and D.D. Breshears, 2010: A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. Forest Ecology and Management, 259(4), 660-684. Almaraz, J.J., F. Mabood, X. Zhou, E.G. Gregorich, and D.L. Smith, 2008: Climate change, weather variability and corn yield at a higher latitude locale: Southwestern Quebec. Climatic Change, 88(2), 187-197. American Society of Civil Engineers, 2009: 2009 Report Card for America's Infrastructure. American Society of Civil Engineers, Reston, Virginia. Anderegg, W.R.L., J.A. Berry, D.D. Smith, J.S. Sperry, L.D.L. Anderegg, and C.B. Field, 2012: The roles of hydraulic and carbon stress in a widespread climate-induced forest die-off. Proceedings of the National Academy of Sciences, 109(1), 233-237. Anderegg, W.R.L., J.M. Kane, and L.D.L. Anderegg, 2013: Consequences of widespread tree mortality triggered by drought and temperature stress. Nature Climate Change, 3, 30-36. Anderson, B.J., H.R. Akcakaya, M.B. Araujo, D.A. Fordham, E. Martinez-Meyer, W. Thuiller, and B.W. Brook, 2009: Dynamics of range margins for metapopulations under climate change. Proceedings of the National Academies of Science, 276(1661), 1415-1420. Anderson, B.G. and M.L. Bell, 2009: Weather-related mortality: how heat, cold, and heat waves affect mortality in the United States. Epidemiology, 20(2), 205. Anderson, C. and S. McLachlan, 2012: Exiting, enduring and innovating: Farm household adaptation to global zoonotic disease. Global Environmental Change, 22, 82-93. Anderson, R.G., J.G. Canadell, J.T. Randerson, R.B. Jackson, B.A. Hungate, D.D. Baldocchi, G.A. Ban-Weiss, G.B. Bonan, K. Caldeira, L. Cao, N.S. Diffenbaugh, K.R. Gurney, L.M. Kueppers, B.E. Law, and S. Luyssaert and T.L. O'Halloran, 2011: Biophysical considerations in forestry for climate protection. Frontiers in Ecology and the Environment, 9, 174-182. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Anderson, M., S. Dobardzic, and D. Gardiner, 2006: Climate change and insurance: An agenda for action in the United States. Allianz Group and WWF, Munchen, Germany and Washington, D.C. Andrachuk, M. and T. Pearce, 2010: Vulnerability and adaptation in two communities in the Inuvialuit settlement region. In: Community Adaptation and Vulnerability in Arctic Regions. [Hovelsrud, G.K. and B. Smit(eds.)]. Springer, New York, pp. 63-81. Anguelovski, I. and J.A. Carmin, 2011: Something borrowed, everything new: innovation and institutionalization in urban climate governance. Current Opinion in Environmental Sustainability, 3(3), 169-175. Arbuckle Jr, J.G., L.S. Prokopy, T. Haigh, J. Hobbs, T. Knoot, C. Knutson, A. Loy, A.S. Mase, J. McGuire, L.W. Morton, J. Tyndall, and M. Widhalm, 2013: Climate change beliefs, concerns, and attitudes toward adaptation and mitigation among farmers in the Midwestern United States. Climate Change, 117, 943-950. Ariano, R., G.W. Canonica, and G. Passalacqua, 2010: Possible role of climate changes in variations in pollen seasons and allergic sensitizations during 27 years. Annals of Allergy, Asthma & Immunology, 104(3), 215-22. Asseng, S., F. Ewert, C. Rosenzweig, J.W. Jones, J.L. Hatfield, A.C. Ruane, K.J. Boote, P.J. Thorburn, R.P. Rotter, D. Cammaranol, N. Brisson, B. Basso, P. Martre, P.K. Aggarwal, C. Angulo, P. Bertuzzi, C. Biernath, A.J. Challinor, J. Doltra, S. Gayler, R. Goldberg, R. Grant, L.N. Kumar, C. Nendel, G. O'Leary, J.E. Olesen, T.M. Osborne, T. Palosuo, E. Priesack, D. Ripoche, M.A. Semenov, I. Shcherbak, P. Steduto, C. Stockle, P. Stratonovitch, T. Streck, I. Supit, F. Tao, M. Travasso, K. Waha, D. Wallach, J.W. White, J.R. Williams, and J. Wolf, 2013: Uncertainty in simulating wheat yields under climate change. Nature Climate Change, . Auffhammer, M. and A. Aroonruengsawat, 2011: Simulating the impacts of climate change, prices and population on California s residential electricity consumption. Climatic Change, 109, S191-S210. Auffhammer, M. and A. Aroonruengsawat, 2012: Erratum to: Simulating the impacts of climate change, prices and population on California s residential electricity consumption. Climatic Change, 113, 1101-1104. Auld, H., J. Waller, S. Eng, J. Klaasen, R. Morris, S. Fernandez, V. Cheng, and D. MacIver, 2010: The Changing Climate and National Building Codes and Standards.Proceedings of American Meteorological Society, Ninth Symposium on the Urban Environment, Session 5.6, August 1-6, 2010, Keystone, CO, US, . Autor, D.H., L.F. Katz, and M.S. Kearney, 2008: Trends in U.S. wage inequality: Revising the revisionists. The Review of Economics and Statistics, 90(2), 300-323. Autorite des Marches Financiers, 2011: Managing Climate Change Risk: Findings of 2010 survey of property and casualty insurers operating in Quebec. AMF. Averyt, K., J. Fisher, A. Huber-Lee, A. Lewis, N. Macknick, J. Madden, S. Rogers, and S. Tellinghuisen, 2011: Freshwater Use by US Power Plants, Union of Concerned Scientists, Cambridge MA. Avise, J., J. Chen, B. Lamb, C. Wiedinmyer, A. Guenther, and E. Salath, 2009: Attribution of projected changes in summertime US ozone and PM 2.5 concentrations to global changes. Atmospheric Chemistry and Physics, 9, 1111-1124. AXA Ipsos, 2012: Individual perceptions of climate change risk. [AXA Ipsos (ed.)], pp. 1-18. Axelson, J.N., D.J. Sauchyn, and J. Barichivich, 2012: New reconstructions of streamflow variability in the South Saskatchewan River Basin from a network of tree ring chronologies, Alberta, Canada. Water Resources Research, 45(9). Azumaya, T., T. Nagasawa, O.S. Temnykh, and G.V. Khen, 2007: Regional and seasonal differences in temperature and salinity limitations of pacific salmon. North Pacific Anadromous Fish Commission, 4, 179-187. Baade, R.A., R. Baumann, and V. Matheson, 2007: Estimating the economic impact of natural and social disasters, with an application to Hurricane Katrina. Urban Studies, 44(11), 2061-2076. Backus, G.A., T.S. Lowry, and D.E. Warren, 2013: The near-term risk of climate uncertainty among the U.S. states. Climatic Change, 116(3-4), 495-522. Badjek, M., E.H. Allison, A.S. Halls, and N.K. Dulvy, 2010: Impacts of climate variability and change on fishery- based livelihoods. Marine Policy, 34, 375-383. Balshi, M.S., A.D. McGuirez, P. Duffy, M. Flannigan, J. Walsh, and J. Melillo, 2009: Assessing the response of area burned to changing climate in western boreal North America using a Multivariate Adaptive Regression Splines (MARS) approach. Global Change Biology, 15, 578-600. Barnett, T.P., J.C. Adam, and D.P. Lettenmaier, 2005: Potential impacts of a warming climate on water availability in snow-dominated regions. Nature, 438(7066), 303-309. Barnett, T.P., D.W. Pierce, H.G. Hidalgo, C. Bonfils, B.D. Santer, T. Das, G. Bala, A. Wood, T. Nozawa, A. Mirin, D. Cavan, and M.D. Dettinger, 2008: Human-induced changes in the hydrology of the western United States. Science, 319(5866), 1080-1083. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Barragan, A.R., L. Bourillon, S. Calme, N. Castelblanco, P. Charruau, J.J. Dominguez Calderon, C. Espadas Manrique, J.M. Espinoza Rodriguez, M. Garcia Rivas, M. Salgado, C. Sotelo, A. Perdomo, W. Hadad, H.H. Arena, V. Hernandez, Y. Gomez, R. Landa, D. Lugo, B. Morales, G. Martinez, O. Moreno, A. Rhodes, Y. Romero, C. Pigueron, C. Pozo, V. Hernandez, I. Salazar, F. Secaira, N. Simones, F. Guerrero, N. Valle, M. Hernandez, and R. Gutierrez, 2011: Adaptación al Cambio Climático en Áreas Protegidas del Caribe de México
. In: Serie Estrategias de Adaptación al Cambio Climático en Áreas Protegidas de México [March, I.J., Cabral, H., Echeverría, Y., Bellot, M. and Frausto, J.M. (eds.)]. Comisión Nacional de Áreas Naturales Protegidas, The Nature Conservancy, Fondo Mexicano para la Conservación de la Naturaleza, México, pp. 1-109. Barsugli, J.J., J.M. Vogel, L. Kaatz, J.B. JSmith, M. Waage, and C. Anderson, 2012: Two faces of uncertainty: Climate science and water utility planning methods. Journal of Water Resources Planning and Management, 138(5), 389-395. Barthel, F. and E. Neumayer, 2012: A trend analysis of normalized insured damage from natural disasters. Climatic Change, 113, 215-237. Battin, J., M.W. Wiley, M.H. Ruckelshaus, R.N. Palmer, E. Korb, K.K. Bartz, and H. Imaki, 2007: Projected impacts of climate change on salmon habitat restoration. Proceedings of the National Academy Sciences, 104(16), 6720-6725. Battisti, D.S. and R.L. Naylor, 2009: Historical warnings of future food insecurity with unprecedented seasonal heat. Science, 323(5911), 240-244. Baumgart-Getz, A., L. Prokopy, and K. Floress, 2012: Why farmers adopt best management practice in the United States: A meta-analysis of the adoption literature. Journal of Environmental Management, 96(1), 17-25. Bee, B., M. Biermann, and P. Tschakert, 2013: Gender, development, and rights-based approaches: Lessons for climate change adaptation and adaptive social protection. Research, Action and Policy: Addressing the Gendered Impacts of Climate Change, , 95-108. Beebe, N.W., R.D. Cooper, P. Mottram, and A.W. Sweeney, 2009: Australia's dengue risk driven by human adaptation to climate change. PLoS Neglected Tropical Diseases, 3(5), e429. Beef Today Editors, 2013: Cargill to idle Texas beef processing plant. AgWeb Farm Journal, January 17. Belanger, D., P. Berry, V. Bouchet, D. Charron, K. Clarke, B. Doyon, M. Fleury, C. Furgal, P. Gosselin, S. Lamy, L.R. Lindsay, G. McBean, N.H. Ogden, J. Seguin, C.J. Shuster, and C.L. Soskolne, 2008a: Human Health in a Changing Climate : A Canadian Assessment of Vulnerabilities and Adaptive Capacity. Health Canada, Ottowa, Ontario, . Belanger, D., B. Peter, V. Bouchet, D. Charron, K. Clarke, B. Doyon, M. Fleury, C. Furgal, P. Gosselin, S. Lamy, L.R. Lindsay, G. McBean, N.H. Ogden, J. Segun, C.J. Shuster, and C.L. Soskolne, 2008b: Human Health in a Changing Climate: A Canadian assessment of Vulnerabilities and Adaptive Capacity. [Segun, J. (ed.)]. Health Canada, Ottawa, ON, pp. 1-484. Bell, M.L., R. Goldberg, C. Hogrefe, P.L. Kinney, K. Knowlton, B. Lynn, J. Rosenthal, C. Rosenzweig, and J.A. Patz, 2007: Climate change, ambient ozone, and health in 50 US cities. Climatic Change, 82(1), 61-76. Bender, M.A., T.R. Knutson, R.E. Tuleya, J.J. Sirutis, G.A. Vecchi, S.T. Garner, and I.M. Held, 2010: Modeled impact of anthropogenic warming on the frequency of intense Atlantic hurricanes. Science, 327(5964), 454-458. Bentz, B.J., J. Régniere, C.F. Fettig, E.M. Hansen, J.L. Hayes, J.A. Hicke, R.G. Kelsey, J.F. Negrón, and S.J. Seybold, 2010: Climate change and bark beetles of the Western United States and Canada: Direct and Indirect Effects. Bioscience, 60(8), 602-613. Berry, H.L., K. Bowen, and T. Kjellstrom, 2010: Climate change and mental health: a causal pathways framework. International Journal of Public Health, 55(2), 123-32. Bethel, J.W., A.N. Foreman, and S.C. Burke, 2011: Disaster preparedness among medically vulnerable populations. American Journal of Preventive Medicine, 40(2), 139-143. Bhatti, J.S., M.J. Apps, and R. Lal, 2006: Anthropogenic changes and the global carbo cycle. In: Climate change and managed ecosystems. [Bhatti, J.S., M.J. Apps, R. Lal, and M.A. Price(eds.)]. Taylor & Francis, Boca Raton, FL, pp. 71-91. Bierbaum, R., J.B. Smith, A. Lee, M. Blair, F.S.I. Chapin, P. Fleming, S. Russo, M. Stults, S. McNeely, E. Wasley, and L. Verduzco, 2012: A comprehensive review of climate adaptation in the United States: more than before, but less than needed. Mitigation and Adaptation Strategies for Global Change, 18(3), 361-401. Bin, O., C. Dumas, B. Poulter, and J. Whitehead, 2007: Measuring the impacts of climate change on North Carolina coastal resources. East Carolina University, Greenville, NC, pp. 101. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bjarndadottir, S., Y. Li, and M.G. Stewart, 2011: A probabilistic-based framework for impact and adaptation assessment of climate change on hurricane damage risks and costs. Structural Safety, 33, 173-185. Black, R., N.W. Adger, N.W. Arnell, S. Dercon, A. Geddes, and D.S.G. Thomas, 2011: The effect of environmental change on human migration. Global Environmental Change, 215, S3-S11. Blake, E.S., C.W. Landsea, and E.J. Gibney, 2007: The deadliest, costliest, and most intense United States tropical cyclones from 1851 to 2006 (and other frequently requested hurricane facts). NOAA/National Weather Service, National Centers for Environmental Prediction, National Hurricane Center, . Blake, R., A. Grimm, T. Ichinose, R. Horton, S. Gaffin, S. Jiong, D. Bader, and D. Cecil, 2011: Urban Climate Processes, trends and projections. In: Climate change and cities: first assessment report of the Urban Climate Change Research Network. [Rosenzweig, C., W.D. Solecki, and S.A. Hammer(eds.)]. Cambridge University Press, New York, NY, pp. 43-83. Bloomberg, M., 2012: Remarks delivered at the New York Marriott Downtown.December 6, 2012, . Bond, W.J. and J.E. Keeley, 2005: Fire as a global herbivore : the ecology and evolution of flammable ecosystems. Trends in Ecology and Evolution, 20, 387-394. Bond, W. and B. Van Wilgen, 1996: Fire and Plants (Population and Community Biology Series 14). Bonsal, B.R., R. Aider, P. Gachon, and S. Lapp, 2012: An assessment of Canadian prairie drought: Past, present, and future. Climate Dynamics, 41(2), 501-516. Bootsma, A., S. Gameda, and D.W. McKenney, 2005: Potential impacts of climate change on corn, soybeans and barley yields in Atlantic Canada. Canadian Journal of Soil Science, 85, 345-357. Bornehag, C.G., G. Blomquist, F. Gyntelberg, B. Jarvholm, P. Malmberg, L. Nordvall, A. Nielsen, G. Pershagen, and J. Sundell, 2001: Dampness in buildings and health. Indoor Air, 11(2), 72-86. Boruff, B.J., C. Emrich, and S.L. Cutter, 2005: Erosion hazard vulnerability of US coastal counties. Journal of Coastal Research, , 932-942. Brenkert-Smith, H., 2010: Building bridges to fight fire: The role of informal social interactions in six Colorado wildland-urban interface communities. International Journal of Wildland Fire, 19, 689-697. Bresch, D. and A. Spiegel, 2011: A blueprint for managing climate risks in emerging markets. Swiss Re. Breshears, D.D., N.S. Cobb, P.M. Rich, K.P. Price, C.D. Allen, R.G. Balice, W.H. Romme, J.H. Kastens, M.L. Floyd, J. Belnap, J.J. Anderson, O.B. Myers, and C.W. Meyer, 2005: Regional vegetation die-off in response to global-change-type drought. Proceedings of the National Academy of Sciences, 102(42), 15144-15148. Bright, E.A., P.R. Coleman, A.N. Rose, and M.L. Urban, 2012: Landscan 2011 Oakridge National Laboratory, Oakridge, TN. British Columbia Ministry of the Environment, 2010: Preparing for Climate Change: British Columbia's Adaptation Strategy. British Columbia Ministry of the Environment, Victoria, British Columbia. Brklacich, M., M. Woodrow, R. McLeman, and K. Vodden, 2008: Enhancing the Capacity of Canadian Rural Communities to Adapt to Uncertain Futures. In: Project No. A1397, June 30. Canadian Climate Impacts and Adaptation Program, Ottawa: Natural Resources Canada. Brooks, S. and M. Loevinsohn, 2011: Shaping agricultural innovation systems responsive to food insecurity and climate change. Natural Resources Forum, 35, 185-200. Brown, P.J., R.S. Bradley, and F.T. Keimig, 2010: Changes in Extreme Climate Indices for the Northeastern United States, 1870-2005. Journal of Climate, 23, 6555-6572. Brown, H.C.P., 2009: Climate change and Ontario forests: Prospects for building institutional adaptive capacity. Mitigation and Adaptation Strategies for Global Change, 14, 513. Browne, M.J. and R.E. Hoyt, 2000: The demand for flood insurance: empirical evidence. Journal of Risk and Uncertainty, 20(3), 291-306. Bryant, C., B. Singh, and P. Thomassin, 2008: Evaluation of agricultural adaptation processes and adaptive capacity to climate change and variability: The co-construction of new adaptation planning tools with stakeholders and farming communities in the Saguenay-Lac-Saint-Jean and Monteregie regions of Quebec. Natural Resources Canada Climate Change Impacts and Adaptation Program. Buechler, S., 2009: Gender, water, and climate change in Sonora, Mexico: implications for policies and programmes on agricultural income-generation. Gender & Development, 17, 51-66. Burch, S., 2010: Transforming barriers into enablers of action on climate change: insights from three municipal case studies in British Columbia, Canada. Global Environmental Change, 20(2), 287-297. Bureau of Reclamation, 2007: Colorado River Interim Guidelines for Lower Basin Shortages and Coordinated Operations for Lake Powell and Lake Mead U.S. Department of the Interior, . Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bureau of Reclamation, 2013: Colorado River Basin Water Supply and Demand Study U.S. Department of the Interior, . Butler, E.E. and P. Huybers, 2012: Adaptation of US maize to temperature variations. Nature Climate Change, 3(1), 68-72. Buttle, J., J.T. Muir, and J. Frain, 2004: Economic impacts of climate change on the Canadian Great Lakes hydro- electric power producers: A supply analysis. Canadian Water Resources Journal, 29, 89-109. Cabas, J., A. Weersink, and E. Olale, 2010: Crop yield response to economic, site and climate variables. Climatic Change, 101, 599-616. Cai, X., D. Wang, and R. Laurent, 2009: Impact of Climate Change on Crop Yield: A Case Study of Rainfed Corn in Central Illinois. Journal of Applied Meteorology & Climatology, 48, 1868-1881. Cakmak, S., R.E. Dales, R.T. Burnett, S. Judek, F. Coates, and J.R. Brook, 2002: Effect of airborne allergens on emergency visits by children for conjunctivitis and rhinitis. Lancet, 359(9310), 947-8. California Department of Water Resources, 2009: Bulletin 160-09 California Water Plan Update 2009. State of California, Sacramento, CA. California Natural Resources Agency, 2009: 2009 California Climate Adaptation Strategy: A Report to the Governor of the State of California in Response to Executive Order S-13-2008. State of California, Sacramento, CA. Campos, M., D. Herrador, C. Manuel, and M.K. McCall, 2013: Adaptation strategies to climate change in two rural communities in Mexico and El Salvador. Boletín De La Asociación De Geógrafos Espanoles, 61, 433-436. Canadian Council of Professional Engineers, 2008: Adapting to Climate Change: Canada's First National Engineering Vulnerability Assessment of Public Infrastructure. Carmin, J., N. Nadkarni, and C. Rhie, 2012: Progress and Challenges in Urban Climate Adaptation Planning: Results of a Global Survey. Massachusetts Institute of Technology, Cambridge, MA. Carriere, A., M. Prevost, A. Zanyadi, P. Chevalier, and B. Barbeau, 2010: Vulnerability of Quebec drinking water treatment plants to cyanotoxins in a climate change context. Journal of Water and Health, 8(3), 455-465. Carroll, M.S., P.J. Cohn, T.B. Paveglio, D.R. Drader, and P.J. Jakes, 2010: Fire burners to firefighters: the Nez Perce and fire. Journal of Forestry, 108(2), 71-76. Cayan, D.R., K.T. Redmond, and L.G. Riddle, 1999: ENSO and hydrologic extremes in the Western United States. Journal of Climate, 12, 2881-2893. Cayan, D.R., T. Das, D.W. Pierce, T.P. Barnett, M. Tyree, and A. Gershunov, 2010: Future dryness in the southwest US and the hydrology of the early 21st century drought. Proceedings of the National Academy of Sciences, 107(50), 21271-21276. CBS Chicago, 2012: City Sets July 5 Record At 103°; Heat Blamed For 2 Deaths CBS News, Chicago Illinois, . Center for Disease Control and Prevention, 2011: Surveillance for Waterborne Disease Outbreaks and Other Health Events Associated with Recreational Water United States, 2007 2008 and Surveillance for Waterborne Disease Outbreaks Associated with Drinking Water United States, 2007 2008. Morbidity and Mortality Weekly Report, 60(12), 1-80. Centre d'expertise hydrique du Québec, 2003: Measures Provided by the Dam Safety Act and Regulation. Government of Quebec, Quebec, Canada. Centre for Research on the Epidemeology of Disasters, 2012: EM-DAT: The International Disaster Data Base Centre for Research on the Epidemeology of Disasters;, Brussels, Belgium, . Chambers, J.Q., J.I. Fisher, H. Zeng, E.L. Chapman, D.B. Baker, and G.C. Hurtt, 2007: Hurricane Katrina's Carbon Footprint on U.S. Gulf Coast Forests. Science, 318(5853), 1107-1107. Chang, W., V.A. Lantz, and D.A. MacLean, 2009: Public attitudes about forest pest outbreaks and control: Case studies in two Canadian provinces. Forest Ecology and Management, 257(4), 1333-1343. Chang, H.H., J. Zhou, and M. Fuentes, 2010: Impact of climate change on ambient ozone level and mortality in southeastern United States. Int. J. Environ. Res. Public Health, 7, 2866-2880. Chen, C.C. and B.A. McCarl, 2009: Hurricanes and possible intensity increases: effects on and reactions from U.S. agriculture. Journal of Agricultural and Applied Economics, 41, 125-144. Chen, F., S. Miao, M. Tewari, J.W. Bao, and H. Kusaka, 2011a: A numerical study of interactions between surface forcing and sea breeze circulations and their effects on stagnation in the greater Houston area. Journal of Geophysical Research: Atmospheres (1984-2012), 116, D12105. Chen, J., W. Xu, J. Velten, Z. Xin, and J. Stout, 2012: Characterization of maize inbred lines for drought and heat tolerance. Journal of Soil and Water Conservation, 67(5), 354-364. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chen, J., F.P. Brissette, and R. Leconte, 2011b: Uncertainty of downscaling method in quantifying the impact of climate change on hydrology. Journal of Hydrology, 401(3), 190-202. Chen, J., J. Avise, B. Lamb, and E. Salath, 2009: The effects of global changes upon regional ozone pollution in the United States. Atmospheric Chemistry and Physics, 9, 1125-1141. Chhetri, N.B., W.E. Easterling, E. Terandoc, and L. Mearns, 2010: Modeling Path Dependence in Agricultural Adaptation to Climate Variability and Change. Annals of the Association of American Geographers, 100(4), 894-907. Chiffoleau, Y., 2009: From Politics to Co-operation: The Dynamics of Embeddedness in Alternative Food Supply Chains. Sociologia Ruralis, 49(3), 218-235. Chinowsky, P., J.C. Price, and J. Neumann, 2013: Assessment of Climate Change Adaptation Costs for the U.S. Road Network. Global Environmental Change, 23, 764-773. Chiotti, Q. and B. Lavender, 2008: Ontario. In: From Impacts to Adaptation: Canada in a Changing Climate. [Lemmen, D.S., F.J. Warren, J. Lacroix, and E. Bush(eds.)]. Government of Canada, Ottawa, pp. 227-274. Choi, W., P.F. Rasmussen, A.R. Moore, and S.J. Kim, 2009: Simulating streamflow response to climate scenarios in central Canada using a simple statistical downscaling method. Climate Research, 40(1), 89-102. Chouinard, O., S. Plante, and G. Martin, 2008: The community engagement process: A governance approach in adaptation to coastal erosion and flooding in Atlantic Canada. Canadian Journal of Regional Science, 31(3), 507-520. Christianson, A., T.K. McGee, and L. L'Hirondelle, 2012a: Community support for wildfire mitigation at Peavine Métis Settlement, Alberta, Canada. Environmental Hazards, 11(3), 177-193. Christianson, A., T.K. McGee, and L. L'Hirondelle, 2012b: How historic and current wildfire experiences in an Aboriginal community influence mitigation preferences. International Journal of Wildland Fire, . City of Vancouver, 2012: Climate Change Adaptation Strategy. City of Vancouver, Vancouver, Canada. Clark, G.E., S.C. Moser, S.J. Ratick, K. Dow, W.B. Meyer, S. Emani, W. Jin, J.X. Kasperson, R.E. Kasperson, and H.E. Schwarz, 1998: Assessing the vulnerability of coastal communities to extreme storms: The case of Revere, MA., USA. Mitigation and Adaptation Strategies for Global Change, 3, 59-82. Clot, B., 2003: Trends in airborne pollen: an overview of 21 years of data in Neuchâtel (Switzerland). Aerobiologia, 19(3), 227-234. Coffee, J.E., J. Parzen, M. Wagstaff, and R.S. Lewis, 2010: Preparing for a changing climate: The Chicago climate action plan's adaptation strategy. Journal of Great Lakes Research, 36, 115-117. Coleman, J.M.A., F.S. Sosa-Rodriguez, L. Mortsch, and P. Deadman, 2013: Assessing Stakeholder Impacts and Adaptation to Low-water Levels: The Trent-Severn Waterway (in review). Climatic Change, . Coles, A.R. and C.A. Scott, 2009: Vulnerability and adaptation to climate change and variability in semi-arid rural southeastern Arizona, USA. Natural Resources Forum, 33(4), 297-309. Collins, M.J., 2009: Evidence for changing flood risk in New England since the late 20th century. Journal of the American Water Resources Association, 45(2), 279-290. Collins, T.W., 2008: The political ecology of hazard vulnerability: marginalization, facilitation and the production of differential risk to urban wildfires in Arizona's White Mountains. Journal of Political Ecology, 15, 21-43. Collins, T.W. and B. Bolin, 2009: Situating hazard vulnerability: people s negotiations with wildfire environments in the US Southwest. Environmental Management, 44(3), 441-455. Collins, T.W., S.E. Grineski, J. Chakraborty, and Y.J. McDonald, 2011: Understanding environmental health inequalities through comparative intracategorical analysis: Racial/ethnic disparities in cancer risks from air toxics in El Paso County, Texas. Health & Place, 17(1), 335-344. Colten, C., R. Kates, and S. Laska, 2008: Community Resilience: Lessons from New Orleans and Hurricane Katrina. In: CAPRI Report 3. Oak Ridge National Laboratory, Oak Ridge, TN. Comfort, L.K., 2006: Cities at Risk. Urban Affairs Review, 41(4), 501. Comisión Económica para América Latina y el Caribe, CEPAL Statistic Databases: Population, By Age Group, by sex Comisión Económica para América Latina y el Caribe, Vitacura, Santiago de Chile. Comisión Intersecretarial de Cambio Climatico, 2009: Programa Especial de Cambio Climatico 2008-2012. Secretaria de Medio Ambiente y Recursos Naturales, Mexico. Comisión Nacional de Áreas Naturales Protegidas and The Nature Conservancy, Programa México, 2009: Programa de manejo integral del fuego, Reserva de la Biosfera Selva El Ocote Chiapas, México 2009- 2012.Mexico, D.F., . Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Comisión Nacional del Agua, 2011: Atlas del Agua en México. Secretaria de Medio Ambiente y Recursos Naturales, México, D.F. Comisión Nacional Forestal, 2012a: Proceedings of Seminario de información estadística y geográfica para prevenir y mitigar los efectos de la sequía e inundaciones en la población y la economía, November 14-15, 2011, Mexico, D.F., . Comisión Nacional para el Conocimiento y Uso de la Biodiversidad, Comisión Nacional de Áreas Naturales Protegidas, The Nature Conservancy, Programa México, and Pronatura, 2007: Análisis de vacíos y omisiones en conservación de la biodiversidad marina de México: océanos, costas e islas. Comisión Nacional para el conocimiento y Uso de la Biodiversidad; Comisión Nacional de Áreas Naturales Protegidas; The Nature Conservancy-Programa México; Pronatura A.C., México, D.F. Committee on Flood Control Alternatives in the American River Basin, Environment and Resources Commission on Geosciences, and Division on Earth and Life Studies and National Research Council, 1995: Flood Risk Management and the American River Basin: An Evaluation. The National Academies Press, Washington, D.C., . Conde, C., R. Ferrer, and S. Orozco, 2006: Climate change and climate variability impacts on rainfed agricultural activities and possible adaptation measures. Atmosfera, 19(3), 181-194. Conrad, E., 2010: Climate Change and Infrastructure in the Gulf of Mexico and Caribbean Basin: New Risks to Building Blocks of Resilience. In: Fighting for Survival: The Vulnerability of America s Gulf Coast and the Caribbean Basin. Tulane University, New Orleans, Louisiana, pp. 63-88. Cooper, D.C. and G. Sehlke, 2012: Sustainability and Energy Development: Influences of Greenhouse Gas Emission Reduction Options on Water Use in Energy Production. Environmental Science & Technology, 46(6), 3509- 3518. Cooper, M.J.P., M.D. Beevers, and M. Oppenheimer, 2008: The potential impacts of sea level rise on the coastal region of New Jersey, USA. Climatic Change, 90, 475-492. Costello, C.J., O. Deschenes, and C.D. Kolstad, 2009: Economic Impacts of Climate Change on California Agriculture
. California Energy Commission, Santa Barbara, CA. Craft, J. and M. Howlett, 2013: Policy Capacity and the Ability to Adapt to Climate Change: Canadian and US Case Studies. Review of Policy Research, 30(1), 1-18. Craine, J.M., A.J. Elmore, K.C. Olson, and D. Tollesons, 2010: Climate change and cattle nutritional stress. Global Change Biology, 16, 2901-2911. Crawford, W., D. Johannessen, F. Whitney, R. Birch, K. Borg, D. Fissel, and S. Vagle, 2007: Appendix C: Physical and Chemical Oceanography in B. G. Lucas, S. Verrin, and R. Brown, editors. In: Ecosystem overview: Pacific North Coast Integrated Management Area (PNCIMA). In: Canadian Technical Report of Fisheries and Aquatic Sciences 2667 [Lucas, B.G., Verrin, S. and Brown, R. (eds.)]. Institute of Ocean Sciences, Sidney, British Columbia. Croci, E., S. Melandri, and T. Molteni, 2010: A Comparative Analysis of Global City Policies in Climate Change Mitigation: London, New York, Milan, Mexico City and Bangkok. In: IEFE Working Paper No. 32. IEFE Bocconi University Working Paper. Cross, M.S., E.S. Zavaleta, D. Bachelet, M.L. Brooks, C.A.F. Enquist, E. Fleishman, L. Graumlich, C.R. Groves, L. Hannah, L. Hansen, G. Hayward, M. Koopman, J. Lawler, J.R. Malcolm, J. Nordgren, B. Petersen, E.L. Rowland, D. Scott, S.L. Shafer, M.R. Shaw, and G.M. Tabor, 2012: The Adaptation for Conservation Targets (ACT) framework: a tool for incorporating climate change into natural resource management. Environmental Management, 50(3), 341-351. Cross, M.S., P.D. McCarthy, G. Garfin, D. Gori, and C.A. Enquist, 2013: Accelerating adaptation of natural resource management to address climate change. Conservation Biology, 27(1), 4-13. Crozier, L.G., R.W. Zabel, and A. Hamlet, 2008: Predicting differential effects of climate change at the population level with life-cycle models of spring Chinook salmon. Global Change Biology, 14(2), 236-249. Cueva-Luna, T.E., R. Few, and A. Mercado, 2011: Afrontando el cambio climático y los riesgos contra la salud: respuestas en la Sierra Tarahumara [Coping with climate change and health risks: responses in the Sierra Tarahumara]. Estudios Demográficos Y Urbanos, 26(3), 671-708. Cuo, L., D.P. Lettenmaier, M. Alberti, and J.E. Richey, 2009: Effects of a century of land cover and climate change on the hydrology of the Puget Sound basin. Hydrological Processes, 23(6), 907-933. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Curriero, F.C., J.a. Patz, J.B. Rose, and S. Lele, 2001: The association between extreme precipitation and waterborne disease outbreaks in the United States, 1948-1994. American Journal of Public Health, 91(8), 1194- 9. Cutter, S., B. Solecki, N. Bragado, J. Carmin, M. Fragkias, M. Ruth, and T. Wilbanks, 2013: Chapter 11 Urban Systems, Infrastructure, and Vulnerability. In: Federal Advisory Committee Draft Climate Assessment. National Climate Assessment and Development Advisory Committee, Washington, D.C., . Cutter, S.L. and C. Emrich, 2005: Are natural hazards and disaster losses in the US increasing? EOS, Transactions American Geophysical Union, 86(41), 381. Cutter, S.L., B.J. Boruff, and W.L. Shirley, 2003: Social Vulnerability to Environmental Hazards*. Social Science Quarterly, 84(2), 242-261. Cutter, S.L., L. Barnes, M. Berry, C. Burton, E. Evans, E. Tate, and J. Webb, 2008: A place-based model for understanding community resilience to natural disasters. Global Environmental Change, 18(4), 598-606. Dai, A.G., 2011: Drought under global warming: a review. Wiley Interdisciplinary Reviews: Climate Change, 2, 45- 65. Daley, M.L., J.D. Potter, and W.H. McDowell, 2009: Salinization of urbanizing New Hampshire streams and groundwater: Effects of road salt and hydrologic variability. Journal of the North American Benthological Society, 28(4), 929-940. Das, T., M.D. Dettinger, D.R. Cayan, and H.G. Hidalgo, 2011: Potential increase in floods in California s Sierra Nevada under future climate projections. Climatic Change, 109(1), 71-94. Davis, J.B., 2004: Healthy Forests Initiatige: Unhealthy Policy Choices in Forest and Fire Management, The. Envtl.L., 34, 1209. Daw, T., W.N. Adger, K. Brown, and M.C. Badjeck, 2009: Climate change and capture fisheries: potential impacts, adaptation and mitigation. In: Climate change implications for fisheries and aquaculture: overview of current scientific knowledge.FAO Fisheries and Aquaculture Technical Paper, No. 530 [Cochrane, K., De Young, C., Soto, D. and Bahri, T. (eds.)]. FAO, Rome, Italy, pp. 7-106. DeGaetano, A.T., 2009: Time-Dependent Changes in Extreme-Precipitation Return-Period Amounts in the Continental United States. Journal of Applied Meteorology and Climatology, 48(10), 2086-2099. Degallier, N., C. Favier, C. Menkes, M. Lengaigne, W.M. Ramalho, R. Souza, J. Servain, and J.P. Boulanger, 2010: Toward an early warning system for dengue prevention: modeling climate impact on dengue transmission. Climatic Change, 98(3), 581-592. Delfino, R.J., 2002: Epidemiologic evidence for asthma and exposure to air toxics: linkages between occupational, indoor, and community air pollution research. Environmental Health Perspectives, 110(Suppl 4), 573. Delpla, I., A.-. Jung, E. Baures, M. Clement, and O. and Thomas, 2009: Impacts of climate change on surface water quality in relation to drinking water production. Environment International, 35(8), 1255-1233. Deschenes, O., M. Greenstone, and J. Guryan, 2009: Climate Change and Birth Weight. American Economic Review, 99(2), 211-217. Desrochers, G., R. Roy, L. Roy, G. Pasher, F. Guay, and D. Tapsoba, 2009: Comparing methods to investigate the impacts of climate change.Proceedings of 2009 AWRA Spring Specialty Conference: Managing Water Resources Development in a Changing Climate, May 4-5, 2009, Anchorage, Alaska, pp. 22. Diffenbaugh, N.S. and C.B. Field, 2013: Changes in ecologically critical terrestrial climate conditions. Science, 341(6145), 486-491. Diffenbaugh, N.S., M. Ashfaq, and M. Scherer, 2011: Transient regional climate change: analysis of the summer climate response in a high-resolution, century-scale, ensemble experiment over the continental United States. Journal of Geophysical Research, 116, D24111. Diffenbaugh, N.S. and M. Scherer, 2011: Observational and model evidence of global emergence of permanent, unprecedented heat in the 20th and 21st centuries. Climatic Change, 107(3-4), 615-624. Diffenbaugh, N.S. and F. Giorgi, 2012: Climate change hotspots in the CMIP5 global climate model ensemble. Climatic Change, 114(3-4), 813-822. Diffenbaugh, N.S., M. Scherer, and M. Ashfaq, 2012: Response of snow-dependent hydrologic extremes to continued global warming. Nature Climate Change, 3(4), 349-384. Direccion General de Comunicacion Social, 2012: Por efectos de la sequía, más de dos millones de Mexicanos están en riesgo de hambruna. Direccion General de Comunicacion Social, . Dirmeyer, P.A., Y. Jin, B. Singh, and X. Yan, 2013: Evolving land-atmosphere interactions over North America from CMIP5 simulations. Journal of Climate, . Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Diuk-Wasser, M., G. Vourc'h, P. Cislo, A.G. Hoen, F. Melton, S.a. Hamer, M. Rowland, R. Cortinas, G.J. Hickling, J.I. Tsao, A.G. Barbour, U. Kitron, J. Piesman, and D. Fish, 2010: Field and climate-based model for predicting the density of host-seeking nymphal Ixodes scapularis , an important vector of tick-borne disease agents in the eastern United States. Global Ecology and Biogeography, 19(4), 504-514. Dolfman, M.L., S. Fortier Wasser, and B. Bergman, 2007: The Effects of Hurricane Katrina on the New Orleans Economy. Monthly Labor Review, 130, 3-18. Dombeck, M.P., J.E. Williams, and C.A. Wood, 2004: Wildfire policy and public lands: Integrating scientific understandings with social concerns across landscapes. Conservation Biology, 18(4), 883-889. Doney, S.C., V.J. Fabry, R.A. Feely, and J.A. Kleypas, 2009: Ocean acidification: The other CO2problem. Annual Review of Marine Science, 1, 169-192. Douglas, E.M., P.H. Kirshen, M. Paolisso, C. Watson, J. Wiggin, A. Enrici, and M. Ruth, 2012: Coastal flooding, climate change and environmental justice: identifying obstacles and incentives for adaptation in two metropolitan Boston Massachusetts communities. Mitigation and Adaptation Strategies for Global Change, 17(5), 537-562. Downing, A. and A. Cuerrier, 2011: A synthesis of the impacts of climate change on the First Nations and Inuit of Canada. Indian Journal of Traditional Knowledge, 10(1), 57-70. Doyle, M.W., E.H. Stanley, D.G. Havlick, M.J. Kaiser, G. Steinbach, W.L. Graf, G.E. Galloway, and J.A. Riggsbee, 2008: Aging infrastructure and ecosystem restoration. Science, 319(5861), 286. Drake, B.G., L. Hughes, E.A. Johnson, B.A. Seibel, M.A. Cochrane, V.J. Fabry, D. Rasse, and L. Hannah, 2005: Synergistic effects. In: Climate Change and Biodiversity. [T.E. Lovejoy and L.J. Hannah (ed.)]. Yale University Press, New Haven, CT, pp. 296-316. Duffy, P. and C. Tebaldi, 2012: Increasing prevalence of extreme summer temperatures in the US. Climatic Change, 111(2), 487-495. Dukes, J.S., N.R. Chiariello, S.R. Loarie, and C.B. Field, 2011: Strong response of an invasive plant species (Centaurea solstitialis L.) to global environmental changes. Ecological Applications, 21(6), 1887-1894. Dumais, D. and M. Prévost, 2007: Management for red spruce conservation in Québec: The importance of some physiological and ecological characteristics-A review. The Forestry Chronicle, 83(3), 378-392. Dupuis, A.P. and B.J. Hann, 2009: Warm spring and summer water temperatures in small eutrophic lakes of the Canadian prairies: Potential implications for phytoplankton and zooplankton. Journal of Plankton Research, 31(5), 489-502. Dwyer, B., B. Harding, E. Wilson, M. Brown, J. Pearce, and J. Smith, 2012: Colorado River Water Availability Study. Colorado Water Conservation Board, Denver, CO, pp. 1-189. Eakin, H., 2006: Weathering risk in rural Mexico: Economic, climatic and institutional change. University of Arizona Press, Tuscon, USA, . Eakin, H. and C. Tucker, 2006: Responding to the coffee crisis: A pilot study of farmers' adaptations in Mexico, Guatemala and Honduras. The Geographical Journal, 172(2), 156-171. Eakin, H. and K. Appendini, 2008: Livelihood change, farming, and managing flood risk in the Lerma valley, Mexico. Agriculture and Human Values, 25(4), 555-566. Eakin, H., A. Winkels, and J. Sendzimir, 2009: Nested vulnerability: exploring cross-scale linkages and vulnerability teleconnections in Mexican and Vietnamese coffee systems. Environmental Science and Policy, 12(4), 398-412. Eakin, H., A.M. Lerner, and F. Murtinho, 2010: Adaptive capacity in evolving peri-urban spaces: Responses to flood risk in the Upper Lerma River Valley, Mexico. Global Environmental Change, 20(1), 14-22. Eakin, H.C. and A. Patt, 2011: Are adaptation studies effective, and what can enhance their practical impact? Wiley Interdisciplinary Review: Climatic Change, 2(2), 141-153. Ebi, K.L., J. Balbus, P.L. Kinney, E. Lipp, D. Mills, M.S. O Neill, and M. Wilson, 2008: Chapter 2: Effects of global change on human health. In: Analyses of the Effects of Global Change on Human Health and Welfare and Human Systems. [Gamble, J.L. (ed.)]. U.S. Environmental Protection Agency, Washington, D.C., pp. 2-1-2- 78. Ebi, K.L., 2011: Resilience to the Health Risks of Extreme Weather Events in a Changing Climate in the United States. International Journal of Environmental Research and Public Health, 8(12), 4582-4595. Emberlin, J., M. Detandt, R. Gehrig, S. Jaeger, N. Nolard, and A. Rantio-Lehtimäki, 2002: Responses in the start of Betula (birch) pollen seasons to recent changes in spring temperatures across Europe. International Journal of Biometeorology, 46(4), 159-170. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Emelko, M.B., U. Silins, K.D. Bladon, and M. and Stone, 2011: Implications of land disturbance on drinking water treatability in a changing climate: Demonstrating the need for "source water supply and protection" strategies. Water Research, 45(2), 461-472. Energy Information Administration, 2013: Annual Energy Outlook 2013. U.S. Energy Information Administration, U.S. Department of Energy, Washington DC. Energy, Mines and Resources: Forest Management Branch, 2012: Forest Health Report 2012. Government of Yukon, Yukon, Canada, pp. 8-15. Environment Canada, 2011a: Federal Adaptation Policy Framework. Government of Canada, Gatineau, QC. Environment Canada, 2011b: Backgrounder: Canada s Ongoing Commitment to Climate Change Adaptation Government of Canada, Ottawa, Canada, pp. 1. Environment Canada, 2013a: Water Availability: Indicator Initiative Government of Canada, Ottawa, Canada, . Environment Canada, 2013b: Regional Freshwater Quality in Canadian Rivers Government of Canada, Ottawan, Canada, . Epstein, P., 2010: The ecology of climate change and infectious diseases: comment. Ecology, 91(3), 925-8; discussion 928-9. Escolero-Fuentes, O.A., S.E. Martínez, S. Kralisch, and M. Perevochtchikova, 2009: Vulnerabilidad de las fuentes de abastecimiento de agua potable de la Ciudad de México en el contexto de cambio climático. Centro Virtual de Cambio Climático, ICyTDF, Instituto de Ciencias de la Atmósfera-UNAM, Instituto de Geología-UNAM, El Colegio de México, Mexico. Esqueda, G.S., J.E. Ospina-Norena, C. Gay-Garcia, and C. Conde, 2010: Vulnerability of water resources to climate change scenarios. Impacts on the irrigation districts in the Guayalejo-Tamesí river basin, Tamaulipas, México. Atmósfera, 24(1), 141-155. Euripidou, E. and V. Murray, 2004: Public health impacts of floods and chemical contamination. Journal of Public Health, 26(4), 376-383. Executive Office of the President, 2013: The President's Climate Action Plan. The White House, Washington DC. Feagin, R.A., D.J. Sherman, and W.E. Grant, 2005: Coastal erosion, global sea-level rise, and the loss of sand dune plant habitats. Frontiers in Ecology and the Environment, 3(7), 359-364. Federal Emergency Management Agency, 2013: Questions about the Biggert-Waters Flood Insurance Reform Act of 2012. Federal Emergency Management Agency. Feltmate, B. and J. Thistlethwaite, 2012: Insurance: Adapting Building Codes for Climate Change. In: Climate Change Adaptation: A Priorities Plan for Canada. University of Waterloo, Waterloo, Ontario, pp. 85-92. Few, R., K. Brown, and E.L. Tompkins, 2007: Public participation and climate change adaptation: avoiding the illusion of inclusion. Climate Policy, 7(1), 46-59. Field, C.B., L.D. Mortsch, M. Brklacich, D.L. Forbes, P. Kovacs, J.A. Patz, S.W. Running, and M.J. Scott, 2007: North America. Climate Change 2007: Impacts, Adaptation and Vulnerability. In: Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Parry, M.L., O.F. Canziani, J.P. Paulikof, van der Linden, P. J., and C.E. Hanson(eds.)]. Cambridge University Press, Cambridge, UK, pp. 617-652. Findell, K.L. and T.L. Delworth, 2010: Impact of Common Sea Surface Temperature Anomalies on Global Drought and Pluvial Frequency. Journal of Climate, 23(3), 485-503. Fischlin, A., G.F. Midgley, J.T. Price, R. Leemans, B. Gopal, C. Turley, M.D.A. Rounsevell, O.P. Dube, J. Tarazona, and A.A. Velichko, 2007: Ecosystems, their properties, goods, and services. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der Linden and C. E. Hanson (ed.)]. Cambridge University Press, Cambridge, pp. 211-272. Flint, C.G., B. McFarlane, and M. Miller, 2008: Human dimensions of forest disturbance by insects: An international synthesis. Environmental Management, 43(6), 1174-1186. Flood, J.F. and L.B. Cahoon, 2011: Risks to Coastal Wastewater Collection Systems from Sea-Level Rise and Climate Change. Journal of Coastal Research, 27(4), 652-660. Follett, R.F. and D.A. Reed, 2010: Soil carbon sequestration in grazing lands: Societal benefits and policy implications. Rangeland Ecology and Management, 63(1), 4-15. Food and Agriculture Organization, 2009: FAOSTAT: Countries by Commodity, Wheat Food and Agriculture Organization of the United Nations, . Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Food and Agriculture Organization, 2012: Forests and Climate Change Working Paper 10. In: Forest Management and Climate Change: a literature review. Food and Agriculture Organization of the United Nations, Rome, pp. 1-55. Forbes, K.A., S.W. Kienzle, C.A. Coburn, J.M. Byrne, and J. Rasmussen, 2011: Simulating the hydrological response to predicted climate change on a watershed in southern Alberta, Canada. Climatic Change, 105(3-4), 555-576. Ford, J.D., T. Pearce, J. Prno, F. Duerden, L.B. Ford, M. Beaumier, and T. Smith, 2010a: Perceptions of climate change risks in primary resource use industries: A survey of the Canadian mining sector. Regional Environmental Change, 10(1), 65. Ford, J.D., T. Pearce, J. Prno, F. Duerden, L.B. Ford, T.R. Smith, and M. Beaumier, 2011: Canary in a coal mine: Perceptions of climate change risks and response options among Canadian mine operations. Climatic Change, 109(3-4), 399-415. Ford, J.D., T. Pearce, F. Duerden, C. Furgal, and B. Smit, 2010b: Climate change policy responses for Canada's Inuit population: The importance of and opportunities for adaptation. Global Environmental Change, 20(1), 177-191. Frank, E., H. Eakin, and D. Lopez-Carr, 2010: Risk perception and adaptation to climate risk in the coffee sector of Chiapas, Mexico.Proceedings of Tropentag 2010: Conference on International Research on Food Security, Natural Resource Management and Rural Development, September 14-16, Zurich, pp. 1-4. Fraser, E.D., E. Simelton, M. Termansen, S.N. Gosling, and A. South, 2013: Vulnerability hotspots : Integrating socio-economic and hydrological models to identify where cereal production may decline in the future due to climate change induced drought. Agricultural and Forest Meteorology, 170, 195-205. Frazier, T.G., N. Wood, B. Yarnal, and D.H. Bauer, 2010: Influence of potential sea level rise on societal vulnerability to hurricane storm-surge hazards, Sarasota County, Florida. Applied Geography, 30(4), 490-505. Frei, T. and E. Gassner, 2008: Climate change and its impact on birch pollen quantities and the start of the pollen season an example from Switzerland for the period 1969 2006. International Journal of Biometeorology, 52(7), 667-674. Fried, J.S., J.K. Gilless, W.J. Riley, T.J. Moody, C. Simon de Blas, K. Hayhoe, M. Moritz, S. Stephens, and M. Torn, 2008: Predicting the effect of climate change on wildfire behavior and initial attack success. Climatic Change, 87, 251-264. Friesinger, S. and P. Bernatchez, 2010: Perceptions of Gulf of St. Lawrence coastal communities confronting environmental change: Hazards and adaptation, Quebec, Canada. Ocean and Coastal Management, 53(11), 669-678. Fuchs, A. and H. Wolff, 2010: Concept and unintended consequences of weather index insurance: the case of Mexico. American Journal of Agricultural Economics, 93(2), 505-511. Furgal, C. and T. Prowse, 2008: Northern Canada. In: From Impacts to Adaptation: Canada in a Changing Climate 2007. [Lemmen, D.S., F.J. Warren, J. Lacroix, and E. Bush(eds.)]. Government of Canada, Ottawa, pp. 57-118. Galindo, I., S. Castro, and M. Valdés, 2009: Satellite derived solar irradiance over Mexico. Atmósfera, 4(3), 189- 201. Galindo, L.M., 2009: The Economics of Climate Change in Mexico. Secretaría de Medio Ambiente y Recursos Naturales y Secretaría de Hacienda y Crédito Público, Mexico City. Gallagher, J., 2012: Learning about an infrequent event: Evidence from flood insurance take-up in the US. Department of Economics, University of California at Berkeley, California, pp. 1-47. Gallivan, F., K. Bailey, and L. O'Rourke, 2009: Planning for Impacts of Climate Change at US Ports. Transportation Research Record: Journal of the Transportation Research Board, 2100(1), 15-21. Gamble, J.L., K.L. Ebi, A.E. Grambsch, F.G. Sussman, and T.J. Wilbanks, 2008: Analyses of the Effects of Global Change on Human Health and Welfare and Human Systems (SAP 4.6). U.S. Environmental Protection Agency, Washington, D. C., pp. 1-283. Gasper, R., A. Blohm, and M. Ruth, 2011: Social and economic impacts of climate change on the urban environment. Current Opinion in Environmental Sustainability, 3(3), 150-157. Gay, C., F. Estrada, C. Conde, H. Eakin, and L. Villers, 2006: Potential impacts of climate change on agriculture: A case of study of coffee production in Veracruz, Mexico. Climatic Change, 79(3-4), 259-288. Gazeau, F., C. Quiblier, J.M. Jansen, J. Gattuso, J.J. Middelburg, and C.H. Heip, 2007: Impact of elevated CO2 on shellfish calcification. Geophysical Research Letters, 34(7). Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Georgakakos, A., P. Fleming, M. Dettinger, C. Peters-Lidgard, T.C. Richmond, K. Reckhow, K. White, and D. Yates, 2013: Chapter 3: Water Resources. In: Federal Advisory Committee Draft Climate Assessment Report Released for Public Review. Georgescu, M., D.B. Lobell, and C.B. Field, 2011: Direct climate effects of perennial bioenergy crops in the United States. Proceedings of the National Academy of Sciences, 108(11), 4307-4312. Getches, D.H., 2003: Water management in the United States and the fate of the Colorado River delta in Mexico. US-Mex.LJ, 11, 107. Girardin, M.P. and D. Sauchyn, 2008: Three centuries of annual area burned variability in northwestern North America inferred from tree rings. The Holocene, 18(2), 205-214. Girardin, M.P., A.A. Ali, C. Carcaillet, M. Mudelsee, I. Drobyshev, C. HELy, and Y. Bergeron, 2009: Heterogeneous response of circumboreal wildfire risk to climate change since the early 1900s. Global Change Biology, 15(11), 2751-2769. Girardin, M.P., A.A. Ali, C. Carcaillet, S. Gauthier, C. Hély, H. Le Goff, A. Terrier, and Y. Bergeron, 2012: Fire in managed forests of eastern Canada: Risks and options. Forest Ecology and Management, . Gleeson, J., P. Gray, A. Douglas, C.J. Lemieux, and G. Nielsen, 2011: A practitioner's guide to climate change adaptation in Ontario's ecosystems. Ontario Centre for Climate Impacts and Adaptation Resources, Sudbury, Ontario, pp. 1-74. Glick, P. and B.A. Stein, 2011: Scanning the conservation horizon: A guide to climate change vulnerability assessment. National Wildlife Federation, Washington, D.C., pp. 1-168. Gómez-álvarez, A., J.L. Valenzuela-García, D. Meza-Figueroa, M. de la O-Villanueva, J. Ramírez-Hernández, and J. Almendariz-Tapia, 2011: Impact of mining activities on sediments in a semi-arid environment: San Pedro river, Sonora, Mexico. Applied Geochemistry, 26(12), 2101-2112. Gomez-Diaz, J.D., A.I. Monterroso-Rivas, J.A.T. Rueda, and M.L.T. Medrano, 2011: Assessing current and potential patterns of 16 forest species driven by climate change scenarios in México. Atmósfera, 24(1), 31-52. Gómez-Mendoza, L. and L. Arriaga, 2007: Modeling the effect of climate change on the distribution of oak and pine species of Mexico. Conservation Biology, 21(6), 1545-1555. Gong, H., A.T. DeGaetano, and L.C. Harrington, 2011: Climate-based models for West Nile Culex mosquito vectors in the Northeastern US. International Journal of Biometeorology, 55(3), 435-436. González Martínez, E., 2006: Plan Rector de la Cafeticultura Nacional. Sistema Producto Cafe Nacional; La Secretaría de Agricultura, Ganadería, Desarrollo Rural, Pesca y Alimentación, México, DF. Goujon, A., W. Lutz, and W. Sanderson, 2004: Future human capital: population projections by level of education. In: The end of world population growth in the 21st century: New challenges for human capital formation and sustainable development. pp. 121-141. Government of Quebec, 2012: Quebec in Action: Greener by 2020: 2013-2020 Government Strategy for Climate Change Adaptation. Government of Quebec, Quebec City, Quebec. Gramig, B.M., J.M. Barnard, and L.S. Prokopy, 2013: Farmer beliefs about climate change and carbon sequestration incentives. Climate Research, 56, 157-167. Greer, A., V. Ng, and D. Fisman, 2008: Climate change and infectious diseases in North America: the road ahead. Canadian Medical Association Journal, 178(6), 715. Grenzeback, L. and A. Luckmann, 2006: Case Study of the Transportation Sector s Response to and Recovery from Hurricanes Katrina and Rita. Transportation Research Board, Washington DC. Groffman, P.M., P. Kareiva, S. Carter, N.B. Grimm, J.J. Lawler, M. Mack, V. Matzek, and H. Tallis, 2013: Chapter 8: Ecosystems, Biodiversity and Ecosystem Services. In: U.S. National Climate Assessment. U.S. Global Change Research Program, Washington, D. C., . Groisman, P.Y., B.G. Sherstyukov, V.N. Razuvaev, R.W. Knight, J.G. Enloe, N.S. Stroumentova, P.H. Whitfield, E. Frland, I. Hannsen-Bauer, and H. Tuomenvirta, 2007: Potential forest fire danger over Northern Eurasia: changes during the 20th century. Global and Planetary Change, 56(3), 371-386. Gude, P., R. Rasker, and J. van den Noort, 2008: Potential for future development on fire-prone lands. Journal of Forestry, 106(4), 198-205. Hajat, S. and T. Kosatsky, 2010: Heat-related mortality: A review and exploration of heterogeneity. Journal of Epidemiology and Community Health, 64(9), 753-760. Halofsky, J.E., D.L. Peterson, K.A. O Halloran, and C. Hawkins-Hoffman, 2011: Adapting to climate change at Olympic National Forest and Olympic National Park. US Department of Agriculture, Forest Service, Pacific Northwest Research Station, . Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hamin, E.M. and N. Gurran, 2009: Urban form and climate change: Balancing adaptation and mitigation in the US and Australia. Habitat International, 33(3), 238-245. Hamlet, A., S.Y. Lee, K. Mickelson, and M. Elsner, 2010: Effects of projected climate change on energy supply and demand in the Pacific Northwest and Washington State. Climatic Change, 102(1-2), 103-128. Hamlet, A.F., 2011: Assessing water resources adaptive capacity to climate change impacts in the Pacific Northwest Region of North America. Hydrology and Earth System Sciences, 15(5), 1427-1443. Hammer, R.B., S.I. Stewart, and V.C. Radeloff, 2009: Demographic trends, the wildland-urban interface, and wildfire management. Society and Natural Resources, 22(8), 777-782. Hanna, E.G., T. Kjellstrom, C. Bennett, and K. Dear, 2011: Climate change and rising heat: Population health implications for working people in Australia. Asia-Pacific Journal of Public Health, 23(2 SUPPL.), 14S-26S. Hannah, L., P.R. Roehrdanz, M. Ikegami, A.V. Shepard, M.R. Shaw, G. Tabor, L. Zhi, P.A. Marquet, and R.J. Hijmans, 2013: Climate change, wine, and conservation. Proceedings of the National Academy of Sciences, 110(17), 6907-6912. Hanson, S., R. Nicholls, N. Ranger, S. Hallegatte, J. Corfee-Morlot, C. Herweijer, and J. Chateau, 2011: A global ranking of port cities with high exposure to climate extremes. Climatic Change, , 1-23. Hardess, L., A. Karst, M. M Lot, S. Morgan, S. Wolfe, R. Matthews, and R. Sydneysmith, 2011: Climate Change and Adaptive Capacity in Aboriginal Communities South of 60 Assessment Report. Centre for Indigenous Environmental Resources, Winnipeg, Manitoba. Hardoy, J. and P. Romero Lankao, 2011: Latin American cities and climate change: challenges and options to mitigation and adaptation responses. Current Opinion in Environmental Sustainability, 3(3), 158-163. Harlan, S.L., A.J. Brazel, L. Prashad, W.L. Stefanov, and L. Larsen, 2006: Neighborhood microclimates and vulnerability to heat stress. Social Science & Medicine, 63(11), 2847-2863. Harlan, S.L., A.J. Brazel, G.D. Jenerette, N.S. Jones, L. Larsen, L. Prashad, and W.L. Stefanov, 2008: In the shade of affluence: The inequitable distribution of the urban heat island. Research in Social Problems and Public Policy, 15, 173-202. Harlan, S.L. and D.M. Ruddell, 2011: Climate change and health in cities: impacts of heat and air pollution and potential co-benefits from mitigation and adaptation. Current Opinion in Environmental Sustainability, 3(3), 126-134. Harper, S.L., V.L. Edge, C. Schuster-Wallace, O. Berke, and S.A. McEwen, 2011: Weather, water quality and infectious gastrointestinal illness in two Inuit communities in Nunatsiavut, Canada: potential implications for climate change. EcoHealth, 8(1), 93-108. Hatfield, J.L., R.M. Cruse, and M.D. Tomer, 2013: Convergence of agricultural intensification and climate change in the Midwestern United States: implications for soil and water conservation. Marine and Freshwater Research, 64(5), 423-435. Hatfield, J.R., K.J. Boote, B.A. Kimball, D.W. Wolfe, D.R. Ort, R.C. Izaurralde, and G.L. Hahn, 2008: Agriculture. In: The Effects of Climate Change on Agriculture, Land Resources, Water Resources, and Biodiversity in the United States. Synthesis and Assessment Product 4.3. [Walsh, M. (ed.)]. U.S. Climate Change Science Program, Washington, D.C., pp. 21-74. Hayhoe, K., D. Cavan, C. Field, P. Frumhoff, E. Maurer, M.N. Miller, S. Moser, S. Schneider, K.N. Cahill, E. Cleland, L. Dale, and J. Verville, 2004: Emissions pathways, climate change and impacts on California. Proceedings of the National Academies of Science, 101(34), 12422-12427. Hayhoe, K., S. Sheridan, L. Kalkstein, and S. Greene, 2010: Climate change, heat waves, and mortality projections for Chicago. Journal of Great Lakes Research, 36, Supplement 2(0), 65-73. Heap, N., 2007: Hot properties: How global warming could transform B.C.'s real estate sector. David Suzuki Foundation, Vancouver. Hebben, T., 2009: Analysis of water quality conditions and trends for the long-term river network: Athabasca River, 1960-2007. Alberta Environment. Hejazi, M.I. and M. Markus, 2009: Impacts of urbanization and climate variability on floods in northeastern Illinois. Journal of Hydrologic Engineering, 14(6), 606-616. Henstra, D., 2012: Toward the Climate-Resilient City: Extreme Weather and Urban Climate Adaptation Policies in Two Canadian Provinces. Journal of Comparative Policy Analysis: Research and Practice, 14(2), 175-194. Hernandez, A., B. Dominguez, P. Cervantes, S. Munoz-Melgarejo, S. Salazar-Lizan, and A. Tejeda-Martinez, 2011: Temperature-humidity index (THI) 1917-2008 and future scenarios of livestock comfort in Veracruz, México. Atmósfera, 24(1), 89-102. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hernández, L., H. Reyes-Bonilla, and E.F. Balart, 2010: Efecto del blanqueamiento del coral por baja temperatura en los crustáceos decápodos asociados a arrecifes del suroeste del golfo de California. Revista Mexicana De Biodiversidad, 81, S113-S119. Hicke, J.A., C.D. Allen, A.R. Desai, M.C. Dietze, R.J. Hall, E.H. (Ted) Hogg, D.M. Kashian, D. Moore, K.F. Raffa, R.N. Sturrock, and J. Vogelmann, 2012: Effects of biotic disturbances on forest carbon cycling in the United States and Canada. Global Change Biology, 18(1), 7-34. Hirsch, R.M. and K.R. Ryberg, 2012: Has the magnitude of floods across the USA changed with global CO2 levels? Hydrological Sciences Journal, 57(1), 1-9. Hoegh-Guldberg, O., L. Hughes, S. McIntyre, D.B. Lindenmayer, C. Parmesan, H.P. Possingham, and C.D. Thomas, 2008: Assisted colonization and rapid climate change. Science, 321(5887), 345-346. Hoerling, M.P., J.K. Eischeid, X.W. Quan, H.F. Diaz, R.S. Webb, R.M. Dole, and D.R. Easterling, 2012: Is a transition to semipermanent drought conditions imminent in the U.S. Great Plains? Journal of Climate, 25(8380), 8386. Hogg, E.H. and P.Y. Bernier, 2005: Climate change impacts on drought-prone forest in western Canada. The Forestry Chronicle, 81(5), 675-682. Hogg, E.H., J.P. Brandt, and M. Michaelian, 2008: Impacts of a regional drought on the productivity, dieback, and biomass of western Canadian aspen forests. Canadian Journal of Forest Research, 38(6), 1373-1384. Holden, Z.A., P. Morgan, M.A. Crimmins, R.K. Steinhorst, and A.M. Smith, 2007: Fire season precipitation variability influences fire extent and severity in a large southwestern wilderness area, United States. Geophysical Research Letters, 34(16), L16708. Holloway, T., S.N. Spak, D. Barker, M. Bretl, C. Moberg, K. Hayhoe, J. Van Dorn, and D. Wuebbles, 2008: Change in ozone air pollution over Chicago associated with global climate change. Journal of Geophysical Research, 113(D22), 1-14. Holmes, J., 2010: The forestry industry. In: What Do We Know? What Do We Need to Know? The State of Canadian Research on Work, Employment and Climate Change. [Lipsig-Mummé, C. (ed.)]. Work in a Warming World Research Programme, York University, Toronto, pp. 149-167. Hori, M.E., D. Nohara, and H.L. Tanaka, 2007: Influence of Arctic Oscillation towards the Northern Hemisphere surface temperature variability under the global warming scenario. Journal of the Meteorological Society of Japan, 85(6), 847-859. Horton, B. and G.R.A. Richardson, 2011: Climate change adaptation and Canadian municipal and business decision-makers: A snapshot of 2009. In: CCIAD Discussion Paper Series. Natural Resources Canada, Ottawa, Canada. Houle, D., A. Bouffard, L. Duchesne, T. Logan, and R. Harvey, 2012: Projections of future soil temperature and water content for three southern Quebec forested sites. Journal of Climate, 25(21), 7690-7701. Howden, S.M., J.F. Soussana, F.N. Tubiello, N. Chhetri, M. Dunlop, and H. Meinke, 2007: Adapting agriculture to climate change. Proceedings of the National Academy of Sciences, 104(50), 19691-19696. Huber-Sannwald, E., F.T. Maestre, J.E. Herrick, and J.F. Reynolds, 2006: Ecohydrological feedbacks and linkages associated with land degradation: a case study from Mexico. Hydrological Processes, 20(15), 3395-3411. Hunt, A. and P. Watkiss, 2011: Climate change impacts and adaptation in cities: a review of the literature. Climatic Change, 104(1), 13-49. Hurd, B.H. and J. Coonrod, 2012: Hydro-economic Consequences of Climate Change in the Upper Rio Grande. Climate Research, 53(2), 103-118. Hurtado-Díaz, M., H. Riojas-Rodríguez, A. Brito-Hernández, and W. Junger, 2011: Temperature, air pollution and pediatric hospitalizations in Mexico City.Proceedings of 23rd Annual Conference of the International Society for Environmental Epidemiology (ISEE) 2011, September 13-16, 2011, . Hurteau, M.D. and M. North, 2010: Carbon recovery rates following different wildfire risk mitigation treatments. Forest Ecology and Management, 260(5), 930-937. Hurteau, M.D., M.T. Stoddard, and P.Z. Fulé, 2011: The carbon costs of mitigating high-severity wildfire in southwestern ponderosa pine. Global Change Biology, 17(4), 1516-1521. Hystad, P.W. and P.C. Keller, 2006: Disaster management: Kelowna tourism industry s preparedness, impact and response to a 2003 major forest fire. Journal of Hospitality and Tourism Management, 13(1), 44-58. Hystad, P.W. and P.C. Keller, 2008: Towards a destination tourism disaster management framework: Long-term lessons from a forest fire disaster. Tourism and Management, 29(1), 151-162. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Institute of Medicine, 2011: Climate Change, the Indoor Environment, and Health. The National Academies Press, Washington, D.C., pp. 286. Instituto Mexicano de Tecnología del Agua, 2010: Efectos del Cambio Climático en los Recursos Hídricos de México, Volumen III. Atlas de Vulnerabilidad Hídrica en México ante el Cambio Climático. Instituto Mexicano de Tecnología del Agua, México. Instituto Nacional de Ecología y Cambio Climático, 2006: Tercera Comunicación Nacional ante la Convención Marco de las Naciones Unidas sobre Cambio Climático (CMNUCC) [Third National Communication to the UN Framework Convention on Climate Change (UNFCCC)]. Secretaria de Medio Ambiente y Recursos Naturales, Mexico, D.F. Instituto Nacional de Ecología y Cambio Climático, 2012a: Quinta Comunicación Nacional ante la Convención Marco de las Naciones Unidas sobre el Cambio Climático. Secretaria de Medio Ambiente y Recursos Naturales, México, D.F. Instituto Nacional de Ecología y Cambio Climático, 2012b: Adaptación al cambio climático en México: visión, elementos y criterios para la toma de decisiones. Secretaría de Medio Ambiente y Recursos Naturales, Mexico, Mexico City. Insurance Institute for Business and Home Safety, 2011: Rating the States: An Assessment of Residential Building Code and Enforcement Systems for Life Safety and Property Protection in Hurricane-Prone Regions. Insurance Institute for Business and Home Safety, Tampa, Florida. Insurance Institute for Business and Home Safety, 2012: Fortified for Safer Living: Builders' Guide. IBHS, Tampa, Florida. International Boundary and Water Commission, 2012: Commission Signs Colorado River Agreement. In: News and Press Releases. International Boundary and Water Commission United States Section, El Paso, Texas. International Council for the Exploration of the Sea, 2011: Report of the Study Group on Designing Marine Protected Area Networks in a Changing Climate (SGMPAN).Proceedings of ICES CM 2011/SSGSUE:01, November 15-19, 2010, Woods Hole, Massachusetts, US, pp. 155. International Institute for Applied System Analysis, 2007: GGI Downscaled Spatially Explicit Socio-Economic Scenario Data Greenhouse Gas Initiative (GGI) Program of the International Institute for Applied Systems Analysis, . Intersecretarial Commission on Climate Change, 2007: National Strategy on Climate Change: Mexico. Government of Mexico, Mexico City. IPCC, 2012: Summary for Policymakers. In:Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. In: A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change [Field, C.B., Barros, V., Stocker, T.F., Qin, D., Dokken, D.J., Ebi, K.L., Mastrandrea, M.D., Mach, K.J., Plattner, G.-., Allen, S.K., Tignor, M. and Midgley, P.M. (eds.)]. Cambridge University Press, Cambridge U.K. and New York, USA. Isaac, M. and D.P. Van Vuuren, 2009: Modeling global residential sector energy demand for heating and air conditioning in the context of climate change. Energy Policy, 37(2), 507-512. Ivers, L.C. and E.T. Ryan, 2006: Infectious diseases of severe weather-related and flood-related natural disasters. Current Opinion in Infectious Diseases, 19(5), 408. Jaakkola, J.J.K., B.F. Hwang, and N. Jaakkola, 2005: Home dampness and molds, parental atopy, and asthma in childhood: a six-year population-based cohort study. Environmental Health Perspectives, 113(3), 357. Jacob, D.J. and D.A. Winner, 2009: Effect of climate change on air quality. Atmospheric Environment, 43(1), 51-63. Jacques, L., C. De Vit, and F. Gagnon-Lebrun, 2010: Status of Climate Change Adaptation in Canada s Agricultural Sector. Government of Canada Policy Research Initiative. Janetos, A., L. Hansen, D. Inouye, B.P. Kelly, L. Meyerson, W. Peterson, and R. Shaw, 2008: Biodiversity. In: The Effects of Climate Change on Agriculture, Land Resources, Water Resources, and Biodiversity in the United States. Synthesis and Assessment Product 4.3 [Walsh, M. (ed.)]. U.S. Climate Change Science Program, Washington, DC, pp. 21-74. Jantarasami, L.C., J.L. Lawler, and C.W. Thomas, 2010: Institutional Barriers to Climate Change Adaptation in U.S. National Parks and Forests. Ecology and Society, 15(4), 33. Jimenez-Moleon, M.C. and M.A. Gomez-Albores, 2011: Waterborne diseases in the state of Mexico, Mexico (2000- 2005). Journal of Water and Health, 9(1), 200-7. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Johansson, M.A., D.A.T. Cummings, and G.E. Glass, 2009: Multiyear climate variability and dengue--El Nino southern oscillation, weather, and dengue incidence in Puerto Rico, Mexico, and Thailand: a longitudinal data analysis. PLoS Medicine, 6(11), e1000168. Johnston, M., T. Williamson, E. Wheaton, V. Wittrock, H. Nelson, H. Hesseln, L. Vandamme, J. Pittman, and M. Lebel, 2008: Limited Report: Climate Change Adaptive Capacity of Forestry Stakeholders in the Boreal Plains Ecozone. In: Climate Change Impacts and Adaptation Program Project A1383. Saskatchewan Research Council, Saskatoon, SK. Johnston, M., T. Williamson, A. Munson, A. Ogden, M. Moroni, R. Parsons, D. Price, and J. Stadt, 2010: Climate change and forest management in Canada: impacts, adaptive capacity and adaptation options. A State of Knowledge report. Sustainable Forest Management Network, Edmonton, Alberta, pp. 16-35. Jones, G.V., M.A. White, O.R. Cooper, and K. Storchmann, 2005: Climate Change and Global Wine Quality. Climatic Change, 73, 319-343. Jonkman, S.N., B. Maaskant, E. Boyd, and M.L. Levitan, 2009: Loss of life caused by the flooding of New Orleans after Hurricane Katrina: analysis of the relationship between flood characteristics and mortality. Risk Analysis, 29(5), 676-698. Jury, M.R., 2008: Climate influence on dengue epidemics in Puerto Rico. International Journal of Environmental Health Research, 18(5), 323-34. Karl, T.R., G.A. Meehl, C.D. Miller, S.J. Hassol, A.M. Waple, and W.L. Murray, 2008: Weather and Climate Extremes in a Changing Climate: Regions of Focus: North America, Hawaii, Caribbean and US Pacific Islands. In: Synthesis and Assessment Product 3.3: Report by the US Climate Change Science Program and the Subcommittee on Global Change Research. U.S. Climate Change Science Program, Washington DC. Keel, B.G. (ed.), 2007: Assisted Migration as a Conservation Strategy for Rapid Climatic Change: Investigating Extended Photoperiod and Mycobiont Distribution forHabenaria repens Nuttall (Orchidaceae) as a Case Study. Diss. PhD, Antioch University, 1-160 pp. Keeley, J.E. and P.H. Zedler, 2009: Large, high-intensity fire events in southern California shrublands: debunking the fine-grain age patch model. Ecological Applications, 19(1), 64-94. Keim, M.E., 2008: Building human resilience: the role of public health preparedness and response as an adaptation to climate change. American Journal of Preventive Medicine, 35(5), 508-516. Kelly, G., 2010: Built to a new code. Canadian Underwriter, 8, 48-50. Kelly, J., P.A. Makar, and D.A. Plummer, 2012: Projections of mid-century summer air quality for North America: Effects of changes in climate and precursor emissions. Atmospheric Chemistry and Physics, 12, 3875-3940. Kelly, A.E. and M.L. Goulden, 2008: Rapid shifts in plant distribution with recent climate change. Proceedings of the National Academy of Sciences, 105(33), 11823-11826. Kemp, A., B.P. Horton, S.J. Culver, D.R. Corbett, O. Van De Plassche, and R. Edwards, 2008: Early Onset of Historical Accelerated Relative Sea Level Rise in North Carolina, USA.Proceedings of Geological Society of America Annual Meeting 2008, 2008, Houston, Texas, . Kenny, G.F., J. Yardley, C. Brown, R.J. Sigal, and O. Jay, 2010: Heat stress in older individuals and patients with common chronic diseases. Canadian Medical Association Journal, 182(10), 1053-1060. Kenny, J.F., N.L. Barber, S.S. Hutson, K.S. Linsey, J.K. Lovelace, and M.A. Maupin, 2009: Estimated use of water in the United States in 2005. US Geological Survey Reston, VA, . Kenter, P., 2010: Climate change adaptation missing. Daily Commercial News and Construction Record, 83(97). Kerkhoven, E. and T.Y. Gan, 2011: Differences and sensitivities in potential hydrologic impact of climate change to regional-scale Athabasca and Fraser River basins of the leeward and windward sides of the Canadian Rocky Mountains respectively. Climatic Change, 106(4), 583-607. Kharin, V.V., F.W. Zwiers, X. Zhang, and M. Wehner, 2013: Changes in temperature and precipitation extremes in the CMIP5 ensemble. Climatic Change, 119(2), 345-357. Kienzle, S.W., M.W. Nemeth, J.M. Byrne, and R.J. MacDonald, 2012: Simulating the hydrological impacts of climate change in the upper North Saskatchewan River basin, Alberta, Canada. Journal of Hydrology, 412, 76- 89. King County Department of Natural Resources and Parks, 2008: Vulnerability of Major Wastewater Facilities to Flooding from Sea-Level Rise. King County (WA) Department of Natural Resources and Parks, Wastewater Treatment Division, Seattle, Washington, pp. 1-13. Kinney, P.L., 2012: Health: A new measure of health effects. Nature Climate Change, 2(4), 233-234. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kinney, P., M. Oneill, M. Bell, and J. Schwartz, 2008: Approaches for estimating effects of climate change on heat- related deaths: challenges and opportunities. Environmental Science & Policy, 11(1), 87-96. Kinney, P., P. Sheffield, R.S. Ostfeld, J. Carr, R. Leichenko, and P. Vancura, 2011: Chapter 11: Public Health. In: Responding to Climate Change in New York State: the ClimAID Integrated Assessment for Effective Climate Change Adaptation in New York State. [Rosenzweig, C., W. Solecki, A. DeGaetano, M. O Grady, and P. Grabhorn(eds.)]. New York State Energy Research and Development Authority, Albany, NY, pp. 397-438. Kinney, P.L., 2008: Climate change, air quality, and human health. American Journal of Preventive Medicine, 35(5), 459-67. Kirchner, J.W., C.M. Austin, A. Myers, and D.C. Whyte, 2011: Quantifying remediation effectiveness under variable external forcing using contaminant rating curves. Environmental Science and Technology, 45(18), 7874-7881. Kirshen, P., M. Ruth, and W. Anderson, 2006: Chapter 7: Climate s Long-term Impacts on Urban Infrastructures and Services: The Case of Metro Boston. In: Regional Climate Change And Variability: Impacts And Responses. [Ruth, M., K. Donaghy, and P.H. Kirshen(eds.)]. Edward Elgar Publishers, Cheltenham, England, pp. 191-225. Kirshen, P., K. Knee, and M. Ruth, 2008: Climate change and coastal flooding in Metro Boston: impacts and adaptation strategies. Climatic Change, 90(4), 453-473. Kjellstrom, T., R.S. Kovats, S.J. Lloyd, T. Holt, and R.S.J. Tol, 2009: The direct impact of climate change on regional labor productivity. Archives of Environmental and Occupational Health, 64(4), 217-227. Kjellstrom, T. and J. Crowe, 2011: Climate change, workplace heat exposure, and occupational health and productivity in central America. International Journal of Occupational and Environmental Health, 17(3), 270- 281. Klerkx, L. and C. Leeuwis, 2009: Establishment and embedding of innovation brokers at different innovation system levels: Insights from the Dutch agricultural sector. Technological Forecasting and Social Change, 76, 849-860. Knowlton, K., M. Rotkin-Ellman, G. King, H.G. Margolis, D. Smith, G. Solomon, R. Trent, and P. English, 2009: The 2006 California heat wave: impacts on hospitalizations and emergency department visits. Environmental Health Perspectives, 117(1), 61-67. Knowlton, K., J.E. Rosenthal, C. Hogrefe, B. Lynn, S. Gaffin, R. Goldberg, C. Rosenzweig, K. Civerolo, J. Ku, and P.L. Kinney, 2004: Assessing Ozone-Related Health Impacts under a Changing Climate. Environmental Health Perspectives, 112(15), 1557-1563. Knutson, T.R., J.L. McBride, J. Chan, K. Emanual, G. Holland, C. Landsea, I. Held, J.P. Kossin, A.K. Srivastava, and M. Sugi, 2010: Tropical cyclones and climate change. Nature Geosciences, 3, 157-163. Koffi, J.K., P.A. Leighton, Y. Pelcat, L. Trudel, L.R. Lindsay, F. Milord, and N.H. Ogden, 2012: Passive surveillance of I. scapularis ticks: Enhanced analysis for early detection of emerging Lyme disease risk. Journal of Medical Entomology, 49(2), 400-409. Kolden, C.A. and T.J. Brown, 2010: Beyond wildfire: perspectives of climate, managed fire, and policy in the USA. International Journal of Wildland Fire, 19, 364-373. Kolivras, K., 2010: Changes in dengue risk potential in Hawaii, USA, due to climate variability and change. Climate Research, 42(1), 1-11. Kovacs, J.M., J. Malczewski, and F. Flores-Verdugo, 2004: Examining local ecological knowledge of hurricane impacts in a mangrove forest using an analytical hierarchy process (AHP) approach. Journal of Coastal Research, 20(3), 792-800. Kovacs, P., 2005: Promoting resilient homes. Canadian Underwriter, May 2005, 36-40. Kovats, R.S. and S. Hajat, 2008: Heat stress and public health: A critical review. Annual Review of Public Health, 29, 41-55. Krishnamurthy, P.K., J.B. Fisher, and C. Johnson, 2011: Mainstreaming local perceptions of hurricane risk into policymaking: A case study of community GIS in Mexico. Global Environmental Change, 21, 143-153. Kucharik, C.J. and S.P. Serbin, 2008: Impacts of recent climate change on Wisconsin corn and soybean yield trends. Environmental Research Letters, 3(3), 034003. Kulshreshtha, S., 2011: Climate change, prairie agriculture, and prairie economy: the new normal. Canadian Journal of Agricultural Economics, 59, 19-44. Kumar, S., V. Merwade, J. Kam, and K. Thurner, 2009: Streamflow trends in Indiana: Effects of long term persistence, precipitation and subsurface drains. Journal of Hydrology, 374(1), 171-183. Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kumar, S., V. Merwade, J.L. Kinter III, and D. Niyogi, 2013a: Evaluation of Temperature and Precipitation Trends and Long-term Persistence in CMIP5 20th Century Climate Simulations. Journal of Climate, . Kumar, S., S.I. Kinter, P.A. Dirmeyer, Z. Pan, and J.M. Adams, 2013b: Multi-decadal Climate Variability and the Warming Hole in North America results from CMIP5 20th and 21st Century Climate Simulations. Journal of Climate, 26(11), 3511-3527. Kundzewicz, Z., L. Mata, N.W. Arnell, P. Döll, B. Jimenez, K. Miller, T. Oki, Z. Sen, and I. Shiklomanov, 2008: The implications of projected climate change for freshwater resources and their management. Hydrological Sciences Journal, 53(1), 3-10. Kunkel, K.E., K.E. Bromirski, H.E. Brooks, T. Cavazos, A.V. Douglas, D.R. Easterling, K.A. Emanuel, P.Y. Groisman, G.J. Holland, T.R. Knutson, J.P. Kossin, P.D. Komar, P.D. Levinson, and R.L. Smith, 2008: Observed Changes in Weather and Climate Extremes in Weather and Climate Extremes in a Changing Climate. Regions of Focus: North America, Hawaii, Caribbean, and U.S. Pacific Islands. [Karl, T.R., Meehl, G.A., Miller, C.D., Hassol, S.J., Waple, A.M. and Murray, W.L. (eds.)]. U.S. Climate Change Science Program and the Subcommittee on Global Change Research, Washington, DC. Kunkel, K.E., H.-. Huang, X.-. Liang, J.-. Lin, D. Wuebbles, Z. Tao, A. Williams, M. Caughey, J. Zhu, and K. Hayhoe, 2007: Sensitivity of future ozone concentrations in the northeast USA to regional climate change. Mitigation and Adaptation Strategies for Global Change, 13(5-6), 597-606. Kunreuther, H.C. and E.O. Michel-Kerjan, 2009: At War with the Weather: Managing Large Scale Risks in a New Era of Catastrophes. Massachusetts Institute of Technology Press, . Kurz, W.A., C.C. Dymond, G. Stinson, G.J. Rampley, E.T. Neilson, A.L. Carroll, T. Ebata, and L. Safranyik, 2008a: Mountain pine beetle and forest carbon feedback to climate change. Nature, 452(24), 987-990. Kurz, W.A., G. Stinson, G.J. Rampley, C.C. Dymond, and E.T. Neilson, 2008b: Risk of natural disturbances makes future contribution of Canada s forests to the global carbon cycle highly uncertain. Proceedings of the National Academy of Sciences, 105(5), 1551-1555. Lafferty, K.D., 2009: The ecology of climate change and infectious diseases. Ecology, 90(4), 888-900. Lal, P., J. Alavalapati, and D.E. Mercer, 2011: Socio-economic impacts of climate change on rural United States. Mitigation and Adaptation Strategies for Global Change, 16, 819-844. Lambrechts, L., K.P. Paaijmans, T. Fansiri, L.B. Carrington, and L.D. Kramer, 2011: Impact of daily temperature fluctuations on dengue virus transmission by Aedes aegypti. Proceedings of the National Academy of Sciences, 108(21), 7460-7465. Laukkonen, J., P.K. Blanco, J. Lenhart, M. Keiner, B. Cavric, and C. Kinuthia-Njenga, 2009: Combining climate change adaptation and mitigation measures at the local level. Habitat International, 33(3), 287-292. Lazo, J.K., M. Lawson, P.H. Larsen, and D.M. Waldman, 2011: U.S. Economic Sensitivity to Weather Variability. Bulletin of the American Meteorological Society, 6, 709-720. Leal Asencio, M.T., D.V. Millán Gómez, C.G. Méndez Jaime, and C.A. Servín Jungdorf, 2008: Evaluación de la afectación de la calidad del agua en cuerpos de agua superficiales y subterráneos por efecto de la variabilidad y el cambio climático y su impacto en la biodiversidad, agricultura, salud, turismo e industria [Assessment of the impacts on surface and groundwater due to climate change and the cliamte variability and effects on biodiversity, agriculture, health tourims and the industry]. Secretaria de Medio Ambiente y Recursos Naturales, Mexico, pp. 123. Lee, J., S. Gryze, and J. Six, 2011: Effect of climate change on field crop production in California's Central Valley. Climatice Change, 109(1), 335-353. Leichenko, R., 2011: Climate change and urban resilience. Current Opinion in Environmental Sustainability, 3(3), 164-168. Leighton, P.A., J.K. Koffi, Y. Pelcat, L.R. Lindsay, and N.H. Ogden, 2012: Predicting the speed of tick invasion: An empirical model of range expansion for the Lyme disease vector Ixodes scapularis in Canada. Journal of Applied Ecology, 49(2), 457-464. Leon, C. and C. Neri, 2010: Case study: Mexico City. In: Climate Change, Disaster Risk Management and the Urban Poor. The World Bank, pp. 217-233. Leonard, L., J. Dorton, S. Culver, and R. Christian, 2009: Coastal and estuarine observing in North Carolina: Integrating observations and science to understand our coastal environment. East Carolina University Press, Carolina, USA, pp. 30. Leurig, S., 2011: Climate Risk Disclosure by Insurers: Evaluating insurer responses to the NAIC climate disclosure survey. Ceres, Boston, MA. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Levetin, E. and d.W. Van, 2008: Changing pollen types/concentrations/distribution in the United States: fact or fiction? Current Allergy and Asthma Reports, 8(5), 418-24. Levina, E., J.S. Jacob, L.E. Ramos, and I. Ortiz, 2007: Policy frameworks for adaptation to climate change in coastal zones: The case of the Gulf of Mexico. Organisation for Economic Co-operation and Development, Paris, France. Liao, K., E. Tagaris, K. Manomaiphiboon, C. Wang, J. Woo, P. Amar, S. He, and A.G. Russell, 2009: Quantification of the impact of climate uncertainty on regional air quality. Atmospheric Chemistry and Physics, 9(3), 865-878. Liao, K., E. Tagaris, K. Manomaiphiboon, S.L. Napelenok, J. Woo, S. He, P. Amar, and A.G. Russell, 2007: Sensitivities of ozone and fine particulate matter formation to emissions under the impact of potential future climate change. Environmental Science & Technology, 41(24), 8355-61. Liceaga-Correa, M.A., L.U. Arellano-Méndez, and H. Hernández--Núnez, 2011: Efectos de los huracanes y cambio climático sobre el Caribe mexicano: adaptabilidad de los pastos marinos. In: Vulnerabilidad de las zonas costeras mexicanas ante el cambio climático. [Botello, A.V., S. Villanueva-Fragoso, J. Gutiérrez, and J.L. Rojas Galaviz(eds.)]. Gobierno del estado de Tabasco, Semarnat-INE, UNAM-ICMyL, Universidad Autónoma de Campeche, Mexico, pp. 211-228. Lin, B.B., 2007: Agroforestry management as an adaptive strategy against potential microclimate extremes in coffee agriculture. Agricultural and Forest Meteorology, 144(1-2), 85-94. Lin, B.B., 2010: The role of agroforestry in reducing water loss through soil evaporation and crop transpiration in coffee agroecosystems. Agricultural and Forest Meteorology, 150, 510-518. Lin, J., K.O. Patten, K. Hayhoe, X. Liang, and D.J. Wuebbles, 2008: Effects of Future Climate and Biogenic Emissions Changes on Surface Ozone over the United States and China. Journal of Applied Meteorology and Climatology, 47(7), 1888-1909. Lin, J., D.J. Wuebbles, H. Huang, Z. Tao, M. Caughey, X. Liang, J. Zhu, and T. Holloway, 2010: Potential effects of climate and emissions changes on surface ozone in the Chicago area. Journal of Great Lakes Research, 36, 59- 64. Liu, J., C. Folberth, H. Yang, J. Röckström, K. Abbaspour, and A.J. Zehnder, 2013: A global and spatially explicit assessment of climate change impacts on crop production and consumptive water use. PLoS ONE, 8(2), e57750. Liu, Y., S. L Goodrick, and J. A Stanturf, 2012: Future US wildfire potential trends projected using a dynamically downscaled climate change scenario. Forest Ecology and Management, . Lo, E. and E. Levetin, 2007: Influence of meteorological conditions on early spring pollen in the Tulsa atmosphere from 1987-2006. Journal of Allergy and Clinical Immunology, 119(1), S101-S101. Loarie, S.R., P.B. Duffy, H. Hamilton, G.P. Asner, C.B. Field, and D.D. Ackerly, 2009: The velocity of climate change. Nature, 462(7276), 1052-1055. Lobell, D.B. and C.B. Field, 2011: California perennial crops in a changing climate. Climate Change, 109(1), 317- 333. Lobell, D.B., U.L.C. Baldos, and T.W. Hertel, 2013: Climate adaptation as mitigation: the case of agricultural investments. Environmental Research Letters, 8(1), 015012. Lobell, D.B., C.B. Field, K.N. Cahill, and C. Bonfils, 2006: Impacts of future climate change on California perennial crop yields: Model projections with climate and crop uncertainties. Agricultural and Forest Meteorology, 141(2), 208-218. Lobell, D.B., S. Wolfram, and J. Costa-Roberts, 2011: Climate Trends and Global Crop Production Since 1980. Science, 333(6042), 616-620. Locke, P., C. Clifton, and S. Westra, 2011: Extreme weather events and the mining industry. Engineering and Mining Journal, 212(3), 58. Lozano-Fuentes, S., M.H. Hayden, C. Welsh-Rodriguez, C. Ochoa-Martinez, B. Tapia-Santos, K.C. Kobylinski, C.K. Uejio, E. Zielinski-Gutierrez, L. Delle Monache, A.J. Monaghan, D.F. Steinhoff, and L. Eisen, 2012: The Dengue Virus Mosquito Vector Aedes aegypti at High Elevation in México. American Journal of Tropical Medicine and Hygiene, 87(5), 902-909. Lund, J., E. Hanak, W. Fleenor, R. Howitt, J. Mount, and P. Moyle, 2007: Envisioning futures for the Sacramento- San Joaquin Delta. Public Policy Institute of California, San Francisco, USA, pp. 285. MacDonald, G.M., 2010: Water, climate change, and sustainability in the southwest. Proceedings of the National Academy of Sciences, 107(50), 21256-21262. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Macias Fauria, M. and E.A. Johnson, 2006: Large-scale climatic patterns control large lightning fire occurrence in Canada and Alaska forest regions. Journal of Geophysical Research, 111(G04008). Macias Fauria, M. and E.A. Johnson, 2008: Climate and wildfires in the North American boreal forest. Philosophical Transactions of the Royal Society B, 363, 2317-2329. MacKendrick, N. and J. Parkins, 2005: Social Dimensions of Community Vulnerability to Mountain Pine Beetle. In: Mountain Pine Beetle Initiative Working Paper 2005-26. Natural Resources Canada, Canadian Forest Service, Pacific Forestry Centre, Victoria, BC. Magana, V., 2010: Climate Change, Disaster Risk Management and the Urban Poor Mexcio City Case Study. In: Estudio de vulnerabilidad diferenciada ante eventos extremos del clima en la zona metropolitan de la ciudad de México: Aspectos climáticos [Leon, C. and C. Neri(eds.)]. World Bank, Washington, D.C., pp. 72-114. Magrin, G., C.G. Garcia, D.C. Choque, J.C. Giménez, A.R. Moreno, G.J. Nagy, C. Nobre, and A. Villamiza, 2007: Chapter 13 - Latin America, IPCC WGII Fourth Assessment Report., pp. 63. Mahmud, a., M. Hixson, J. Hu, Z. Zhao, S.-. Chen, and M.J. Kleeman, 2010: Climate impact on airborne particulate matter concentrations in California using seven year analysis periods. Atmospheric Chemistry and Physics, 10(22), 11097-11114. Mailhot, A. and S. Duchesne, 2010: Design Criteria of Urban Drainage Infrastructures under Climate Change. Journal of Water Resources Planning and Management, 136(2), 201-208. Mansourian, S., A. Belokurov, and P.J. Stephenson, 2009: The role of forest protected areas in adaptation to climate change. Unasylva, 60(1-2), 63-68. Mantua, N., I. Tohver, and A. Hamlet, 2010: Climate change impacts on streamflow extremes and summertime stream temperature and their possible consequences for freshwater salmon habitat in Washington State. Climatic Change, 102(1-2), 187-223. Marlin, A., L. Olsen, D. Bruce, J. Ollerhead, K. Singh, J. Heckman, B. Walters, D. Meadus, and A. Hanson, 2007: Examining Community Adaptive Capacity to Address Climate Change, Sea Level Rise, and Salt Marsh Restoration in Maritime Canada. Rural and Small Town Program, Sackville, NB. Marshall, E. and T. Randhir, 2008: Effect of climate change on watershed system: A regional analysis. Climatic Change, 89(3-4), 263-280. Marshall, G.R., 2013: Transaction costs, collective action and adaptation in managing complex social ecological systems. Ecological Economics, 88, 185-194. Marshall, N.A., S.E. Park, W.N. Adger, K. Brown, and S.M. Howden, 2012: Transformational capacity and the influence of place and identity. Environmental Research Letters, 7(3), 034022. Martiello, M.A. and M.V. Giacchi, 2010: Review Article: High temperatures and health outcomes: A review of the literature. Scandinavian Journal of Public Health, 38(8), 826-837. Martin, W.E., I.M. Martin, and B. Kent, 2009: The role of risk perceptions in the risk mitigation process: The case of wildfire in high risk communities. Journal of Environmental Management, 91, 489-498. Martinez Arroyo, A., S. Manzanilla Naim, and J. Zavala Hidalgo, 2011: Vulnerability to climate change of marine and coastal fisheries in México. Atmosfera, 24(1), 103-123. Maryland Commission on Climate Change Adaptation and Response Working Group, 2008: Comprehensive Strategy for Reducing Maryland's Vulnerability to Climate Change: Phase I: Sea-level rise and coastal storms. Maryland Department of Natural Resources, Annapolis, Maryland. Maryland Department of the Environment on behalf of the Maryland Commission on Climate Change, 2010: Update to Governor and General Assembly. Maryland Department of the Environment, Annapolis, Maryland. Matthews, S.N., L.R. Iverson, A.M. Prasad, and M.P. Peters, 2011: Changes in potential habitat of 147 North American breeding bird species in response to redistribution of trees and climate following predicted climate change. Ecography, 260, 1460-1472. Mawdsley, J.R., R. O Malley, and D.S. Ojima, 2009: A review of climate change adaptation strategies for wildlife management and biodiversity conservation. Conservation Biology, 23(5), 1080-1089. Mazari-Hiriart, M., Y. Lopez-Vidal, S. Ponce-de-Leon, J.J. Calva, F. Rojo-Callejas, and C.-. Gonzalo, 2005: Longitudinal study of microbial diversity and seasonality in the Mexico City metropolitan area water supply system. Applied and Environmental Microbiology, 71(9), 5129-5137. McBoyle, G.D., Scott, and B. Jones, 2007: Climate change and the future of snowmobiling in non-mountainous regions of Canada. Managing Leisure, 12, 237-250. McDonald, T., 2012: Coastal Impacts, Adaptation, and Vulnerabilities: A Technical Input to the 2013 National Climate Assessment. Island Press, . Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McFarlane, B.L., 2006: Human dimensions of fire management in the wildland-urban interface: A literature review. In: Canadian Wildland Fire Strategy: Background, Syntheses, Analyses, and Perspectives. [Kirsch, K.G. and Fuglem, P. (eds.)]. Canadian Council for Forest Ministries, Edmonton, Alberta, CA, pp. 27-34. McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks of climate change and human settlements in low elevation coastal zones. Environment and Urbanization, 19(1), 17. McGregor, G.R., 2011: Human biometeorology. Progress in Physical Geography, 36(1), 93-109. McKee, K.L., 2011: Biophysical controls on accretion and elevation change in Caribbean mangrove ecosystems. Estuarine, Coastal and Shelf Science, 91, 475-483. McKenzie, D., Z. Gedalof, D.L. Peterson, and P. Mote, 2004: Climatic change, wildfire, and conservation. Conservation Biology, 18(4), 890-920. McLeman, R. and G. Gilbert, 2008: Economic and social adaptation to climate change in Canadian seasonal- economy communities. In: Final Scientific Report for Natural Resources Canada, Project A1319. Natural Resources Canada, Ottawa, Canada. McLeman, R.A., 2010: Impacts of population change on vulnerability and the capacity to adapt to climate change and variability: a typology based on lessons from a hard country . Population & Environment, 31(5), 286-316. McMichael, A.J., P. Wilkinson, R.S. Kovats, S. Pattenden, S. Hajat, B. Armstrong, N. Vajanapoom, E.M. Niciu, H. Mahomed, and C. Kingkeow, 2008: International study of temperature, heat and urban mortality: the ISOTHURM project. International Journal of Epidemiology, 37(5), 1121-1131. Medina-Ramon, M. and J. Schwartz, 2007: Temperature, Temperature Extremes, and Mortality: A Study of Acclimatization and Effect Modification in 50 United States Cities. Occupational and Environmental Medicine, 64(12), 827-833. Meehl, G.A., C. Tebaldi, G. Walton, D. Easterling, and L. McDaniel, 2009: Relative increase of record high maximum temperatures compared to record low minimum temperatures in the U.S. Geophysical Research Letters, 36(23), L23701. Meehl, G.A., J.M. Arblaster, and G. Branstator, 2012: Mechanisms contributing to the warming hole and the consequent US east-west differential of heat extremes. Journal of Climate, 25(18), 6394 6408. Megdal, S.B. and C.A. Scott, 2011: The Importance of Institutional Asymmetries to the Development of Binational Aquifer Assessment Programs: The Arizona-Sonora Experience. Water, 3(3), 949-963. Mehta, V.K., V.R. Haden, B.A. Joyce, D.R. Purkey, and L.E. Jackson, 2013: Irrigation demand and supply, given projections of climate and land-use change, in Yolo County, California. Agricultural Water Management, 117, 70-82. Mercer, K.L., H.R. Perales, and J.D. Wainwright, 2012: Climate change and the transgenic adaptation strategy: Smallholder livelihoods, climate justice, and maize landraces in Mexico. Global Environmental Change, 22(2), 495-504. Meyer, M.D., E. Rowan, C. Snow, and A. Choate, 2013: Impacts of Extreme Weather on Transportation: National Symposium Summary. American Association of State Highway and Transportation Officials, Washington, D.C., pp. 20. Meza-Figueroa, D., R.M. Maier, M. de la O-Villanueva, A. Gómez-Alvarez, A. Moreno-Zazueta, and J. Rivera, 2009: The impact of unconfined mine tailings in residential areas from a mining town in a semi-arid environment: Nacozari, Sonora, Mexico. Chemosphere, 77(1), 140-147. Miao, S., F. Chen, Q. Li, and S. Fan, 2011: Impacts of Urban Processes and Urbanization on Summer Precipitation: A Case Study of Heavy Rainfall in Beijing on 1 Aug 2006. Journal of Applied Meteorology and Climatology, 50, 806-825. Michaelian, M., E.H. Hogg, R.J. Hall, and E. Arsenault, 2011: Massive mortality of aspen following severe drought along the southern edge of the Canadian boreal forest. Global Change Biology, 17(6), 2084-2094. Millar, C.I., N.L. Stephenson, and S.S. Stephens, 2007: Climate change and forests of the future: managing in the face of uncertainty. Ecological Applications, 17(8), 2145-2151. Miller, A.W., A.C. Reynolds, C. Sobrino, and G.F. Riedel, 2009: Shellfish face uncertain future in high CO2 world: influence of acidification on oyster larvae calcification and growth in estuaries. PLoS One, 4(5), e5661. Millerd, F., 2011: The potential impact of climate change on Great Lakes international shipping. Climatic Change, 104(3-4), 629-652. Mills, B.N., S.L. Tighe, J. Andrey, J.T. Smith, and K. Huen, 2009: Climate Change Implications for Flexible Pavement Design and Performance in Southern Canada. Journal of Transportation Engineering, 135(10), 773- 782. Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mills, E. and E. Lecomte, 2006: From Risk to Opportunity: How Insurers can proactively and profitably manage climate change. Ceres, Boston, MA. Mills, E., 2007: From Risk to Opportunity 2007: Insurer responses to climate change. Ceres. Mills, E., 2009: A global review of insurance industry responses to climate change. Geneva Papers on Risk & Insurance, 34(3), 323-359. Mills, E., 2012: The greening of insurance. Science, 338, 1424-1425. Minero, F.J.G., P. Candau, J. Morales, and C. Tomas, 1998: Forecasting olive crop production based on ten consecutive years of monitoring airborne pollen in Andalusia (southern Spain). Agriculture, Ecosystems & Environment, 69(3), 201-215. Ministry of Municipal Affairs and Housing, 2011: Potential changes for the next edition of the Building Code: Second Round of Consultation (February-April 2011). Queen's Printer for Ontario, Toronto, pp. 46. Minville, M., F. Brissette, S. Krau, and R. Leconte, 2009: Adaptation to climate change in the management of a Canadian water-resources system exploited for hydropower. Water Resources Management, 23(14), 2965-6986. Mirza, M.S. and M. Haider, 2003: The State of Infrastructure in Canada: Implications for Infrastructure Planning and Policy. McGill University, Montreal. Molnar, J.J., 2010: Climate change and societal response: Livelihoods, communities, and the environment. Rural Sociology, 75(1), 1-16. Monkkonen, P., 2011: The Housing Transition in Mexico: Expanding Access to Housing Finance. Urban Affairs Review, 47(5), 672-695. Montero Martinez, J., J. Martinez, N.I. Castillo, and B.E. Espinoza, 2010: Capitulo 2. Escenarios climáticos en México proyectados para el siglo XXI: precipitación y temperaturas máxima y mínima [Chapter 2. Climate scenarios projected for the twenty-first century in Mexico: precipitation and maximum and minimum temperatures]. In: Atlas de vulnerabilidad hídrica en México ante el cambio climático [Atlas of water vulnerability to climate change in Mexico]. Secretaría del Medio Ambiente y Recursos Naturales, Mexico, D.F., pp. 41-63. Monterroso Rivas, A.I., C.C. Alvarez, G.R. Dorantes, and J.D.G. Diaz, 2011: Assessing current and potential rainfed maize suitability under climate change scenarios in Mexico. Atmosfera, 24(1), 53-67. Monterroso, A., C. Conde, C. Gay, D. Gómez, and J. López, 2012: Two methods to assess vulnerability to climate change in the Mexican agricultural sector. Mitigation and Adaptation Strategies for Global Change, , 1-17. Moran, D., A. Lucas, and A. Barnes, 2013: Commentary: Mitigation win-win. Nature Climate Change, 3, 611-613. Morello-Frosch, R., M. Pastor Jr, C. Porras, and J. Sadd, 2002: Environmental Health Perspectives, 110(Suppl 2), 149. Morin, C.W. and A.C. Comrie, 2010: Modeled response of the West Nile virus vector Culex quinquefasciatus to changing climate using the dynamic mosquito simulation model. International Journal of Biometeorology, 54(5), 517-29. Moritz, C., C. Hoskin, J. MacKenzie, B. Phillips, M. Tonione, N. Silva, J. VanDerWal, S. Williams, and C. Graham, 2009: Identification and dynamics of a cryptic suture zone in tropical rainforest. Proceedings of the Royal Society B: Biological Sciences, 276(1660), 1235-1244. Morton, T., P. Bretschneider, D. Coley, and T. Kershaw, 2011: Building a better future: An exploration of beliefs about climate change and perceived need for adaptation within the building industry. Building and Environment, 46, 1151-1158. Mortsch, L., J. Bhatti, R. Lal, M. Apps, and M. Price, 2006: Impact of climate change on agriculture, forestry and wetlands. In: Taylor & Francis, CRC Press, Boca Raton, FL, pp. 45-67. Moser, C. and D. Satterthwaite, 2008: Towards pro-poor adaptation to climate change in the urban centres of low- and middle-income countries. World Bank. Moser, S.C. and J.A. Ekstrom, 2010: A framework to diagnose barriers to climate change adaptation. Proceedings of the National Academy of Sciences, 107(51), 22026-22031. Moser, S.C., 2013: Navigating the political and emotional terrain of adaptation: Community engagement when climate change comes home. In: Successful Adaptation to Climate Change: Linking Science and Practice in a Rapidly Changing World. [Moser, S.C. and M.T. and Boykoff(eds.)]. Routledge, London, pp. 289. Mote, P.W., 2006: Climate-driven variability and trends in mountain snowpack in western North America. Journal of Climate, 19, 6209-6220. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mumby, P.J., I.A. Elliott, C.M. Eakin, W. Skirving, C.B. Paris, H.J. . Edwards, S. Enríquez, R. Iglesias-Prieto, L.M. Cherubin, and J.R. Stevens, 2011: Reserve design for uncertain responses of coral reefs to climate change. Ecology Letters, 14, 132-140. Mumme, S.P., 1999: Managing acute water scarcity on the U.S.-Mexico border: Institutional issues raised by the 1990's drought. Natural Resources Journal, 39(1), 149-166. Munich Re, 2011: Topics Geo Natural Catastrophes 2010. In: Topics Geo. Munich Re. Munich Re, 2012: Severe Weather in North America: Perils, Risks, Insurance. In: Knowledge Series: Natural Hazards. Munich Re, Munich, Germany. Murazaki, K. and P. Hess, 2006: How does climate change contribute to surface ozone change over the United States? Journal of Geophysical Research, 111, 1-16. Nadler, A.J. and P.R. Bullock, 2011: Long-term changes in heat and moisture related to corn production on the Canadian Prairies. Climatic Change, 104, 339-352. Naeher, L.P., M. Brauer, M. Lipsett, J.T. Zelikoff, C.D. Simpson, J.Q. Koenig, and K.R. Smith, 2007: Woodsmoke health effects: a review. Inhalation Toxicology, 19(1), 67-106. Najjar, R.G., C.R. Pyke, M.B. Adams, D. Breitburg, C. Hershner, M. Kemp, R. Howarth, M.R. Mulholland, M. Paolisso, D. Secor, K. Sellner, D. Wardrop, and R. Wood, 2010: Potential climate-change impacts on the Chesapeake Bay. Estuarine, Coastal and Shelf Science, 86, 1-20. Nakaegawa, T., A. Kitoh, and M. Hosaka, 2013: Discharge of major global rivers in the late 21st century climate projected with the high horizontal resolution MRI-AGCMs. Hydrological Processes, . Nakashima, D.J., K. Galloway McLean, H.D. Thulstrup, A. Ramos Castillo, and J.T. Rubis, 2011: Proceedings of the Expert Workshop on Indigenous Peoples, Marginalized Populations and Climate Change.Proceedings of Indigenous Peoples, Marginalized Populations and Climate Change, July 19-21, 2011, Mexico City, Mexico, . NatCatSERVICE, 2010: Data source for Chart 24.4 Population projections, children and seniors Statistics Canada, Canada. National Climate Assessment Development Advisory Committee, 2013: Draft National Climate Assessment Report. U.S. Global Change Research Program, Washington D.C. National Climate Data Center, 2013: State of the Climate: Wildfires for Annual 2012 U.S. National Oceanographic and Atmospheric Administration, National Climate Data Center, Asheville NC, . National Energy Board, 2011: Canada's Energy Future: Energy Supply and Demand Projections to 2035. National Energy Board, Ottawa. National Oceanic and Atmospheric Administration, 2010: Adapting to Climate Change: A Planning Guide for State Coastal Managers. NOAA Office of Ocean and Coastal Resource Management, Silver Spring, MD, pp. 1-132. National Oceanic and Atmospheric Administration, 2013: Service Assessment: Hurricane/Post-Tropical Cyclone Sandy October 22 29, 2012. U.S. Department of Commerce, Silver Spring, Maryland. National Research Council, 2010: America s Climate Choices: Panel on Adapting to the Impact of Climate Change. The National Academies Press, Washington D.C. National Round Table on the Environment and the Economy, 2012: Facing the Elements: Building Business Resilience in a Changing Climate (Case Studies). National Round Table on the Environment and the Economy, Ottawa. National Water Commission of Mexico, 2010: Statistics on Water in Mexico, 2010 edition. [Ministry of Environment and Natural Resources (ed.)], Mexico. Natural Resources Canada, 2008: From impacts to adaptation: Canada in a changing climate 2007. [Lemmen, D.S., Warren, F.J., Lacroix, J. and Bush, E. (eds.)]. Government of Canada, Ottawa, ON, pp. 119-170. Natural Resources Canada, 2011: Canada s Regional Adaptation Collaborative Program - Helping Canadians prepare for and adapt to climate change. Government of Canada, Ottawa, Canada. Natural Resources Canada, 2013: Adaptation Platform Natural Resources Canada, Ottawa, . Nawrotski, R.J., F. Riosmena, and L.M. Hunter, 2013: Do rainfall deficits predict U.S.-bound migration from rural Mexico? Evidence from the Mexican Census. Population Research Policy Review, 32, 129-158. Nelson, G.C., M.W. Rosegrant, J. Koo, R. Robertson, T. Sulser, T. Zhu, C. Ringler, S. Msangi, A. Palazzo, M. Batka, M. Magalhaes, R. Valmonte-Santos, M. Ewing, and D. Lee, 2009: Climate Change: Impact on Agriculture and Costs of Adaptation. International Food Policy Research Institute, Washington, DC, USA. Neria, Y. and J.M. Shultz, 2012: Mental health effects of hurricane Sandy: characteristics, potential aftermath, and response. JAMA: The Journal of the American Medical Association, 308(24), 2571. Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 New York City Department of Environmental Protection, 2008: Columbia University Center for Climate Systems Research and HydroQual Environmental Engineers and Scientists. Columbia University Center for Climate Systems Research and HydroQual Environmental Engineers and Scientists, New York, NY. New York State Climate Action Council, 2012: Climate Action Plan Interim Report. New York State Climate Action Council. Newland, K., D.R. Agunias, and A. Terrazas, 2008: Learning by doing: experiences of circular migration. In: Insight: Programs on Migrants, Migration, and Development. Migration Policy Institute, Washington, D.C., pp. 1-28. Newswire, 2011: Thai flooding impact spreads across world for Toyota. Reuters, . Nicholls, R., S. Hanson, C. Herweijer, N. Patmore, S. Hallegatte, J. Corfee-Morlot, J. Château, and R. Muir-Wood, 2008: Ranking port cities with high exposure and vulnerability to climate extremes. In: OECD Environment Working Papers, No. 1. OECD Publishing, pp. 1-62. Nielsen-Gammon, J.W., 2011: The 2011 Texas Drought. A Briefing Packet for the Texas Legislature. The Office of the State Climatologist Report, Texas A&M University, College Station. Niven, R.V., N.M. Hutson, L. Loftus-Otway, and L.B. Boske, 2010: Transportation Infrastructure Planning in Mexico.Proceedings of Transportation Research Board 89th Annual Meeting, . Nolte, C.G., A.B. Gilliland, C. Hogrefe, and L.J. Mickley, 2008: Linking global to regional models to assess future climate impacts on surface ozone levels in the United States. Journal of Geophysical Research, 113, 1-14. Nordstrom, D.K., 2009: Acid rock drainage and climate change. Journal of Geochemical Exploration, 100(2-3), 97- 104. North, M., B.M. Collins, and S. Stephens, 2012: Using fire to increase the scale, benefits, and future maintenance of fuels treatments. Journal of Forestry, 110(7), 392-401. Ntelekos, A.A., M. Oppenheimer, J.A. Smith, and A.J. Miller, 2010: Urbanization, climate change and flood policy in the united states. . Climatic Change, 103(3), 597-616. Nunavut Department of Sustainable Development, 2003: Nunavut Climate Change Strategy. Nunavut Department of Sustainable Development, Iqaluit, Nunavut. Obeysekera, J., M. Irizarry, J. Park, J. Barnes, and T. Dessalegne, 2011: Climate change and its implications for water resources management in south florida. . Stochastic Environmental Research and Risk Assessment, 25(4), 495-516. Ogden, A.E. and J.L. Innes, 2009: Application of structured decision making to an assessment of climate change vulnerabilities and adaptation options for sustainable forest management. Ecology and Society, 14(1), 11. Ogden, N.H., L. St-Onge, I.K. Barker, S. Brazeau, M. Bigras-Poulin, D.F. Charron, C.M. Francis, A. Heagy, L.R. Lindsay, A. Maarouf, P. Michel, F. Milord, C. O'Callaghan J., L. Trudel, and R.A. Thompson, 2008: Risk maps for range expansion of the Lyme disease vector, Ixodes scapularis, in Canada now and with climate change. International Journal of Health Geographics, 7, 24-24. Okey, T.A., H.M. Alidina, A. Montenegro, V. Lo, and S. Jessen, 2012: Climate change impacts and vulnerabilities in Canada s Pacific marine ecosystems. World Wildlife Fund/Canadian Parks and Wilderness Society, Vancouver, Vancouver, BC, pp. 1-157. O'Neill, M.S. and K.L. Ebi, 2009: Temperature extremes and health: impacts of climate variability and change in the United States. Journal of Occupational and Environmental Medicine, 51(1), 13. Ontario Ministry of Environment, 2011: Climate Ready: Ontario's Adaptation Strategy and Action Plan 2011-2014. Queen's Printer for Ontario, pp. 124. Oppenheimer, M., 2013: Climate change impacts: accounting for the human response. Climatic Change, 117, 439- 449. Organisation for Economic Co-operation and Development, 2010: Chapter 1. General Assessment of the Macroeconomic Situation. In: OECD Economic Outlook Volume 2010/2. OECD Publishing, pp. 1-344. Orlowsky, B. and S.I. Senevirantne, 2013: Elusive drought: Uncertainty in observed trends and short- and long-term CMIP5 projections. Hydrology and Earth System Sciences, 17, 1765-1781. Ouellet, C., D. Saint-Laurent, and F. Normand, 2012: Flood events and flood risk assessment in relation to climate and land-use changes: Saint-François River, southern Québec, Canada. Hydrological Sciences Journal, 57(2), 313-325. Overpeck, J. and B. Udall, 2010: Dry Times Ahead. Science, 328(5986), 1642-1643. Pan, Z., R.W. Arritt, E.S. Takle, W.J.J. Gutowski, C.J. Anderson, and M. Segal, 2004: Altered hydrologic feedback in a warming climate introduces a ''warming hole''. Geophysical Research Letters, 31(17), L17109. Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Park, W.A., C.D. Allen, A.K. Macalady, D. Griffin, C.A. Woodhouse, D.M. Meko, T.W. Swetnam, S.A. Rauscher, R. Seager, H. Grissino-Mayer, J.S. Dean, E.R. Cook, C. Gangodagamage, M. Cai, and N.G. McDowell, 2013: Temperature as a potent driver of regional forest drought stress and tree mortality. Nature Climate Change, 3, 292-297. Parkins, J.R. and N.A. MacKendrick, 2007: Assessing community vulnerability: A study of the mountain pine beetle outbreak in British Columbia, Canada. Global Environmental Change, 17, 460-471. Parkins, J.R., 2008: The metagovernance of climate change: Institutional adaptation to the mountain pine beetle epidemic in British Columbia. Journal of Rural and Community Development, 3(2), 7-26. Parmesan, C., 2006: Ecological and evolutionary responses to recent climate change. Annual Review of Ecology Evolution and Systematics, 37, 637-669. Parris, A., C. Simpson, and S. Abdelrahim, 2010: RISA Workshop Report: Looking Ahead at Climate Service, Assessment, and Adaptation. National Oceanographic and Atmospheric Administration, Climate Program Office, Silver Spring, Maryland. Patriquin, M.N., A.M. Wellstead, and W.A. White, 2007: Beetles, trees and people: Regional economic impact sensitivity and policy considerations related to the mountain pine beetle infestation in British Columbia, Canada. Forest Policy and Economics, 9, 938-946. Patz, J., S.J. Vavrus, C.K. Uejio, and S.L. McLellan, 2008: Climate Change and Waterborne Disease Risk in the Great Lakes Region of the U.S. American Journal of Preventive Medicine, 35(5), 451-458. Paudel, K.P. and L.U. Hatch, 2012: Global warming, impact on agriculture and adaptation strategy. Natural Resource Modeling, 25(3), 456-481. Peach, J. and J. Williams, 2007: Population and Economic Dynamics on the U.S.-Mexican Border: Past, Present, and Future. In: The U.S.-Mexican Border Environment: A Road Map to a Sustainable 2020. [Ganster, P. (ed.)]. SCERP and SDSU Press, San Diego, CA, . Pearce, T.D., J.D. Ford, J. Prno, F. Duerden, J. Pittman, M. Beaumier, L. Berrang-Ford, and B. Smit, 2011: Climate change and mining in Canada. Mitigation and Adaptation Strategies for Global Change, 16(3), 347-368. Pearson, C.J., D. Bucknell, and G.P. Laughlin, 2008: Modelling crop productivity and variability for policy and impacts of climate change in eastern Canada. Environmental Modelling & Software, 23, 1345-1355. Peng, C., Z. Ma, X. Lei, Q. Zhu, H. Chen, W. Wang, S. Liu, W. Li, X. Fang, and X. Zhou, 2011: A drought-induced pervasive increase in tree mortality across Canada's boreal forests. 1(9), 467-471. Perkins, B., D. Ojima, and R. Corell, 2007: A survey of climate change adaptation planning. The H. John Heinz III Center for Science, Economics and the Environment, pp. 1-52. Peter, B., S. Want, T. Mogus, and B. Wilson, 2006: Fire risk and population trends in Canada s wildland-urban interface. In: Canadian Wildland Fire Strategy: Background, Syntheses, Analyses, and Perspectives [Kirsch, K.G. and Fuglem, P. (eds.)]. Canadian Council for Forest Ministries, Edmonton, Alberta, CA, pp. 37-48. Peterson, T.C., R.R. Heim Jr, R. Hirsch, D.P. Kaiser, H. Brooks, N.S. Diffenbaugh, R.M. Dole, J.P. Giovannettone, K. Guirguis, T.R. Karl, R.W. Katz, K. Kunkel, D. Lettenmair, G.J. McCabe, C.J. Paciorek, K.R. Ryberg, S. Schubert, V.B.S. Silva, B.G. Stewart, A.V. Vecchia, G. Villiarini, R.S. Vose, J. Walsh, M. Wehner, D. Wolock, K. Wolter, G.A. Woodhouse, and D. Wuebbles, 2013: Monitoring and understanding changes in heat waves, cold waves, floods and droughts in the United States: State of Knowledge. Bulletin of the American Meteorological Society, 94(6), 821-834. Peterson, T.C. and M.O. Baringer, 2009: State of the Climate in 2008. Bulletin of the American Meteorological Society, 90, S1-S196. Peterson, T.C., X. Zhang, M. Brunet-India, and J.L. Vázquez-Aguirre, 2008: Changes in North American extremes derived from daily weather data. Journal of Geophysical Research, 113, 7113. Phillips, O.L., L. Aragao, S.L. Lewis, J.B. Fisher, J. Lloyd, G. Lopez-Gonzalez, Y. Malhi, A. Monteagudo, J. Peacock, and C.A. Quesada, 2009: Drought sensitivity of the Amazon rainforest. Science, 323(5919), 1344- 1347. Picketts, I.M., J. Curry, and E. Rapaport, 2012: Community Adaptation to Climate Change: Environmental Planners' Knowledge and Experiences in British Columbia, Canada. Journal of Environmental Policy & Planning, 14(2), 119-137. Pielke Jr, R.A., J. Gratz, C.W. Landsea, D. Collins, M.A. Saunders, and R. Musulin, 2008: Normalized hurricane damage in the United States: 1900-2005. Natural Hazards Review, 9(1), 29. Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pierce, J. and G. Meyer, 2008: Long-term fire history from alluvial fan sediments: the role of drought and climate variability, and implications for management of Rocky Mountain forests. International Journal of Wildland Fire, 17(1), 84-95. Pinkerton, K.E., W.N. Rom, M. Akpinar-Elci, J.R. Balmes, H. Bayram, O. Brandli, J.W. Hollingsworth, P.L. Kinney, H.G. Margolis, W.J. Martin, E.N. Sasser, K.R. Smith, and T.K. Takaro, 2012: An Official American Thoracic Society Workshop Report: Climate Change and Human Health. Proceedings of the American Thoracic Society, 9(1), 3-8. Poiani, K.A., R.L. Goldman, J. Hobson, J.M. Hoekstra, and K.S. Nelson, 2011: Redesigning biodiversity conservation projects for climate change: examples from the field. Biodiversity and Conservation, 20(1), 185- 201. Poirier, B. and R.C. de Loë, 2012: Protecting aquatic ecosystems in heavily-allocated river systems: the case of the Oldman River Basin, Alberta. The Canadian Geographer, 55(2), 243-261. Ponce-Reyes, R., E. Nicholson, P.W. Baxter, R.A. Fuller, and H. Possingham, 2013: Extinction risk in cloud forest fragments under climate change and habitat loss. Diversity and Distributions, 19(5-6), 518-529. Population Division, Department of Economic and Social Affairs, 2011: World Population Prospects: The 2010 Revision and World Urbanization Prospects United Nations, . Porter, J.R. and M.A. Semenov, 2005: Crop responses to climatic variation. Philosophical Transactions of the Royal Society B-Biological Sciences, 360, 2021-2035. Portmann, R.W., S. Solomon, and G.C. Hegerl, 2009: Spatial and seasonal patterns in climate change, temperatures, and precipitation across the United States. Proceedings of the National Academy of Sciences, 106, 7324. Posey, J., 2009: The determinants of vulnerability and adaptive capacity at the municipal level: Evidence from floodplain management programs in the United States. Global Environmental Change, 19(4), 482-493. Potera, C., 2011: Indoor air quality: Climate change impacts indoor environment. Environmental Health Perspectives, 119(9), A382. Powell, T., D. Hanfling, and L. Gostin, 2012: Emergency preparedness and public health: the lessons of Hurricane Sandy. JAMA Online, , 12-189. Praskievicz, S. and H. Chang, 2011: Impacts of climate change and urban development on water resources in the Tualatin River Basin, Oregon. Annals of the Association of American Geographers, 101(2), 249-271. Pratchett, M.S., P.L. Munday, S.K. Wilson, N.A.J. Graham, J.E. Cinner, D.R. Bellwood, G.P. Jones, N.V.C. Polunin, and T.R. McClanahan, 2008: Effects of climate-induced coral bleaching on coral-reef fishes-ecological and economic consequences. Oceanography and Marine Biology: An Annual Review, 46, 251-296. Preston, B.L., R.M. Westaway, and E.J. Yuen, 2011: Climate Adaptation Planning in Practice: an Evaluation of Adaptation Plans from Three Developed Nations. Mitigation and Adaptation Strategies for Global Change, 16(4), 407. Prieto-González, R., V.E. Cortés-Hernández, and M.J. Montero-Martínez, 2011: Variability of the standardized precipitation index over México under the A2 climate change scenario. Atmosfera, 24(3), 243-249. Prokoph, A., J. Adamowski, and K. Adamowski, 2012: Influence of the 11year solar cycle on annual streamflow maxima in Southern Canada. Journal of Hydrology, 442, 55-62. Pryor, S.C., R.J. Barthelmie, D.T. Young, E.S. Takle, R.W. Arritt, D. Flory, W.J.J. Gutowski, A. Nunes, and J. Roads, 2009: Wind speed trends over the contiguous United States. Journal of Geophysical Research: Atmospheres (1984-2012), 114(D14), D14105. Pye, H.O.T., H. Liao, S. Wu, L.J. Mickley, D.J. Jacob, D.K. Henze, and J.H. Seinfeld, 2009: Effect of changes in climate and emissions on future sulfate-nitrate-ammonium aerosol levels in the United States. Journal of Geophysical Research, 114, 1-18. Quist, D. and I.H. Chapela, 2001: Transgenic DNA introgressed into traditional maize landraces in Oaxaca, Mexico. Nature, 414(6863), 541-543. Racherla, P.N. and P.J. Adams, 2009: U.S. ozone air quality under changing climate and anthropogenic emissions. Environmental Science & Technology, 43(3), 571-7. Radeloff, V.C., R.B. Hammer, S.I. Stewart, J.S. Fried, S.S. Holcomb, and J.S. McKeefry, 2005: The Wildland- Urban Interface in the United States. Ecological Applications, 15(3), 799-805. Raffa, K.F., B.H. Aukema, B.J. Bentz, A.L. Carroll, J.A. Hicke, M.G. Turner, and W.H. Romme, 2008: Cross-scale Drivers of Natural Disturbances Prone to Anthropogenic Amplification: The Dynamics of Bark Beetle Eruptions. Bioscience, 58(6), 501-517. Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ramos, M.M., H. Mohammed, E. Zielinski-Gutierrez, M.H. Hayden, J.L.R. Lopez, M. Fournier, A.R. Trujillo, R. Burton, J.M. Brunkard, L. Anaya-Lopez, A.A. Banicki, P.K. Morales, B. Smith, J. Munoz L., and S.H. Waterman, 2008: Epidemic dengue and dengue hemorrhagic fever at the Texas-Mexico border: results of a household-based seroepidemiologic survey, December 2005. The American Journal of Tropical Medicine and Hygiene, 78(3), 364-9. Rasmussen, A., 2002: The effects of climate change on the birch pollen season in Denmark. Aerobiologia, 18(3), 253-265. Ray, A., J. Vogel, J.B. Smith, and P. Kovacs, 2013: Adaptation to Climate Change in the United States: Examples from Municipalities, Water Resources, and Insurance. Stratus Consulting, Boulder, CO. Reba, M.L., D. Marks, A. Winstral, T.E. Link, and M. Kumar, 2011: Sensitivity of the snowcover energetics in a mountain basin to variations in climate. Hydrological Processes, 25(21), 3312-3321. Reisen, F. and S.K. Brown, 2009: Australian firefighters' exposure to air toxics during bushfire burns of autumn 2005 and 2006. Environment International, 35(2), 342-352. Reisen, F., D. Hansen, and C. Meyer, 2011: Exposure to bushfire smoke during prescribed burns and wildfires: Firefighters exposure risks and options. Environment International, 37(2), 314-321. Reiss, N.M. and S.R. Kostic, 1976: Pollen season severity and meteorologic parameters in central New Jersey. The Journal of Allergy and Clinical Immunology, 57(6), 609-14. Reiter, P., 2008: Climate change and mosquito-borne disease : knowing the horse before hitching the cart. Revue Scientifique Et Technique - Office International Des Epizooties, 27(2), 383-398. Repetto, R., 2008: The Climate Crisis and the Adaptation Myth. In: Working Paper Number 13 [Coppock, J. (ed.)]. Yale School of Forestry and Environmental Studies, pp. 1-20. Reuters, 2012: USACE Fighting to Keep Mississippi River Open. gCaptain, . Richardson, G.R.A., 2010: Adapting to Climate Change: An Introduction for Canadian Municipalities. Government of Canada, Natural Resources Canada, Ottawa, pp. 1-40. Riojas-Rodriguez, H., M. Hurtado-Diaz, G.L. Moreno-Banda, and A. Castaneda Martinez, 2011: Mapping Human Health Vulnerability to Climate Change in México. In: Environmental health perspectives Proceedings of 23rd Annual Conference of the International Society for Environmental Epidemiology (ISEE) 2011, Sept 13-16, 2011, Barcelona, Spain, . Rittmaster, R., W. Adamowicz, B. Amiro, and R.T. Pelletier, 2006: Economic analysis of health effects from forest fires. Canadian Journal of Forest Research, 36(4), 868-877. Romero-Lankao, P., 2010: Water in Mexico City: what will climate change bring to its history of water-related hazards and vulnerabilities? Environment and Urbanization, 22(1), 157. Romero-Lankao, P., 2007: How do local governments in Mexico City manage global warming? Local Environment, 12(5), 519-535. Romero-Lankao, P., 2012: Governing Carbon and Climate in the Cities: An Overview of Policy and Planning Challenges and Options. European Planning Studies, 20(1), 7-26. Romero-Lankao, P., H. Qin, S. Hughes, Borbor-Cordova M., and M. and Haeffner, 2012a: Urban Vulnerability and Adaptation to the Health Impacts of Air Pollution and Climate Extremes in Latin American Cites. In: Urban Areas and Global Climate Change Research in Urban Sociology. Emerald Group Publishing, pp. 247-274. Romero-Lankao, P., S. Hughes, A. Rosas-Huerta, R. Borquez, and D. Gnatz, 2013a: Urban Institutional Response Capacity for Climate Change: An examination of construction and pathways in Mexico City and Santiago. Environment and Planning C, (In press). Romero-Lankao, P. and D. Dodman, 2011: Cities in transition: transforming urban centers from hotbeds of GHG emissions and vulnerability to seedbeds of sustainability and resilience: Introduction and Editorial overview. Current Opinion in Environmental Sustainability, 3(3), 113-120. Romero-Lankao, P. and H. Qin, 2011: Conceptualizing urban vulnerability to global climate and environmental change. Current Opinion in Environmental Sustainability, 3(3), 142-149. Romero-Lankao, P., H. Qin, and K. Dickinson, 2012b: Urban vulnerability to temperature-related hazards: A meta- analysis and meta-knowledge approach. Global Environmental Change, 22(3), 670-683. Romero-Lankao, P., H. Qin, and M. Borbor-Cordova, 2013b: Exploration of health risks related to air pollution and temperature in three Latin American cities. Social Science & Medicine, 83, 110-118. Romsdahl, R., L. Atkinson, and J. Schultz, 2013: Planning for climate change across the US Great Plains: concerns and insights from government decision-makers. Journal of Environmental Studies and Sciences, 3, 1-14. Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rood, S.B., J. Pan, K.M. Gill, C.G. Franks, G.M. Samuelson, and A. Shepherd, 2008: Declining summer flows of Rocky Mountain rivers: Changing seasonal hydrology and probable impacts on floodplain forests. Journal of Hydrology, 349(3), 397-410. Roorda, M.J., A. Páez, C. Morency, R. Mercado, and S. Farber, 2010: Trip generation of vulnerable populations in three Canadian cities: A spatial ordered probit approach. Transportation, 37(3), 525-548. Root, T.L., D.P. MacMynowski, M.D. Mastrandrea, and S.H. Schneider, 2005: Human-modified temperatures induce species changes: joint attribution. Proceedings of the National Academy of Sciences, 102(21), 7465- 7469. Rose, S. and R. Shaw, 2008: The gamble: Circular Mexican migration and the return on remittances. Mexican Studies/Estudios Mexicanos, 24(1), 79-111. Rose, J.B., P.R. Epstein, E.K. Lipp, B.H. Sherman, S.M. Bernard, and J.a. Patz, 2001: Climate variability and change in the United States: potential impacts on water- and foodborne diseases caused by microbiologic agents. Environmental Health Perspectives, 109 Suppl, 211-21. Rosenthal, J., 2009: Climate change and the geographic distribution of infectious diseases. EcoHealth, 6(4), 489-95. Rosenzweig, 2007: Assessment of observed changes and responses in natural and managed systems. In: Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der Linden, and C. E. Hanson (ed.)]. Cambridge University Press, Cambridge, UK, pp. 79- 131. Rosenzweig, C., D.C. Major, K. Demong, C. Stanton, R. Horton, and M. Stults, 2007: Managing climate change risks in New York City s water system: assessment and adaptation planning. Mitigation and Adaptation Strategies for Global Change, 12(8), 1391-1409. Rosenzweig, C. and W. Solecki, 2010: Introduction to climate change adaptation in New York City: building a risk management response. Annals of the New York Academy of Sciences, 1196(1), 13-18. Rosenzweig, C., W. Solecki, S.A. Hammer, and S. Mehrotra, 2010: Cities lead the way in climate-change action. Nature, 467(7318), 909-911. Rosenzweig, C., W.D. Solecki, R. Blake, M. Bowman, C. Faris, V. Gornitz, R. Horton, K. Jacob, A. LeBlanc, and R. Leichenko, 2011: Developing coastal adaptation to climate change in the New York City infrastructure-shed: process, approach, tools, and strategies. Climatic Change, 106(1), 93-127. Ruddell, D., S.L. Harlan, S. Grossman-Clarke, and G. Chowell, 2011: Scales of perception: public awareness of regional and neighborhood climates. Climatic Change, 111(3-4), 581-607. Ruggiero, P., C.A. Brown, P.D. Komar, J.C. Allan, D.A. Reusser, and S. Rumril, 2010: Chapter 6. Impacts of climate change on Oregon's coasts and estuaries. In: Oregon Climate Assessment Report. [Dello, K.D. and P.W. Mote(eds.)]. College of Oceanic and Atmospheric Sciences, Oregon State University, Corvallis, OR, pp. 15-17. Ruiz-Moreno, D., I.S. Vargas, K.E. Olson, and L.C. Harrington, 2012: Modeling Dynamic Introduction of Chikungunya Virus in the United States. PLoS Neglected Tropical Diseases, 6(11), e1918. Russell, R.C., 2009: Mosquito-borne disease and climate change in Australia: time for a reality check. Australian Journal of Entomology, 48(1), 1-7. Sahoo, G., S.G. Schladow, J.E. Reuter, and R. Coats, 2011: Effects of climate change on thermal properties of lakes and reservoirs, and possible implications. Stochastic Environmental Research and Risk Assessment, 25(4), 445- 456. Sakurai, G., T. Iizumi, and M. Yokozawa, 2011: Varying temporal and spatial effects of climate on maize and soybean affect yield prediction. Climate Research, 49, 143-154. Saldana-Zorilla, S. and S. Sandberg, 2009: Impact of climate-related disasters on human migration in Mexico: A spatial model. Climatic Change, 96, 97-118. Saldana-Zorrilla, S.O., 2008: Stakeholders views in reducing rural vulnerability to natural disasters in Southern Mexico: Hazard exposure and coping and adaptive capacity. Global Environmental Change, 18, 583-597. Sallenger, A.H., Jr., K.S. Doran, and P.A. Howd, 2012: Hotspot of accelerated sea-level rise on the Atlantic coast of North America. Nature Climate Change, 2, 884-888. Sánchez-Cortés, M.S. and E.L. Chavero, 2011: Indigenous perception of changes in climate variability and its relationship with agriculture in a Zoque community of Chiapas, Mexico. Climatic Change, 107(3-4), 363-389. Sandel, B., L. Arge, B. Dalsgaard, R. Davies, K. Gaston, W. Sutherland, and J. Svenning, 2011: The influence of Late Quaternary climate-change velocity on species endemism. Science, 334(6056), 660-664. Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sander-Regier, R., R. McLeman, M. Brklacich, and M. Woodrow, 2009: Planning for climate change in Canadian rural and resource-based communities. Environments Journal, 37(1), 35-57. Sarah, B., 2010: Transforming barriers into enablers of action on climate change: Insights from three municipal case studies in British Columbia, Canada. Global Environmental Change, 20(2), 287-297. Satterthwaite, D., S. Huq, H. Reid, M. Pelling, and P. Romero-Lankao, 2007: Adapting to Climate Change in Urban Areas - The possibilities and constraints in low- and middle-income nations. In: Human Settlements Discussion Paper Series. International Institute for Environment and Development, London, UK. Savonis, M.J., V.R. Burkett, and J.R. Potter, 2008: Impacts of Climate Change and Variability on Transportation Systems and Infrastructure: Gulf Coast Study, Phase I. In: U.S. Climate Change Science Program Synthesis and Assessment Product 4.7. U.S. Climate Change Science Program, Washington DC. Schaeffer, R., A.S. Szklo, Pereira de Lucena, A. F., Moreira Cesar Borba, B. S., L.P. Pupo Nogueira, F.P. Fleming, A. Troccoli, M. Harrison, and M.S. Boulahya, 2012: Energy sector vulnerability to climate change: A review. Energy, 38(1), 1-12. Scheckenberger, R., A. Farrell, and M. Senior, 2009: Economic Assessment of Climate Change Scenarios on Drainage Infrastructure Design.Proceedings of 2009 Annual Conference and Exhibition of the Transportation Association of Canada-Transportation in a Climate of Change, Vancouver, BC, . Schindler, D.W. and W.F. Donahue, 2006: An impending water crisis in Canada's western prairie provinces. Proceedings of the National Academy of Sciences, 103(19), 7210-7216. Schlenker, W., W.M. Hanemann, and A.C. Fisher, 2007: Water Availability, Degree Days, and the Potential Impact of Climate Change on Irrigated Agriculture in California. Climatic Change, 81, 19-38. Schlenker, W. and M.J. Roberts, 2009: Nonlinear temperature effects indicate severe damages to US crop yields under climate change. Proceedings of the National Academy of Sciences, 106(37), 15594-15598. Schloss, C.A., T.A. Nunez, and J.J. Lawler, 2012: Dispersal will limit ability of mammals to track climate change in the Western Hemisphere. Proceedings of the National Academy of Sciences, 109(22), 8606-8611. Scholze, M., W. Knorr, N.W. Arnell, and I.C. Prentice, 2006: A climate-change risk analysis for world ecosystems. Proceedings of the National Academy of Sciences, 103(35), 13116-13120. Schroth, G., P. Laderach, J. Dempewolf, S.M. Philpott, J. Haggar, H. Eakin, T. Castillejos, J. Garcia Moreno, L. Soto Pinto, R. Hernandez, A. Eitzinger, and J. Ramirez-Villegas, 2009: Towards a climate change adaptation strategy for coffee communities and ecosystems in the Sierra Madre de Chiapas, Mexico. Mitigation and Adaptation Strategies for Global Change, 14(7), 605-625. Schwalm, C.R., C.A. Williams, K. Schaefer, D. Baldocchi, T.A. Black, A.H. Goldstein, B.E. Law, W.C. Oechel, K.T.P. U, and R.L. Scott, 2012: Reduction in carbon uptake during turn of htecentury drought in western North America. Nature Geoscience, 5, 551-555. Scott, D. and G. McBoyle, 2007: Climate change adaptation in the ski industry. Mitigation and Adaptation Strategies for Global Change, 12(8), 1411-1431. Scott, D., B. Amelung, S. Becken, J. Ceron, G. Dubois, S. Gössling, P. Peeters, and M. Simpson, 2008a: Technical Report. In: Climate Change and Tourism: Responding to Global Challenges. In: Climate Change and Tourism: Responding to Global Challenges. World Tourism Organization and United Nations Environment Programme, Madrid, Spain, pp. 23-250. Scott, D.J., Dawson, and B. Jones, 2008b: Climate change vulnerability of the US Northeast winter recreation- tourism sector. Mitigation and Adaptation Strategies for Global Change, 13, 577-596. Scott, D.G., McBoyle, and A. Minogue, 2007: Climate change and quebec's ski industry. Global Environmental Change, 17, 181-190. Scott, J., 2007: Agricultural policy and rural poverty in Mexico. Working Paper. CIDE, Mexico, D.F., . Seager, R., M. Ting, I. Held, Y. Kushnir, J. Lu, G. Vecchi, H. Huang, N. Harnik, A. Leetmaa, N. Lau, C. Li, J. Velez, and N. Naik, 2007: Model projections of an imminent transition to a more arid climate in southwestern North America. Science, 316(5828), 1181-1184. Seager, R. and G.A. Vecchi, 2010: Greenhouse warming and the 21st century hydroclimate of southwestern North America. Proceedings of the National Academy of Sciences, 107(50), 21277-21282. Seager, R., A. Tzanova, and J. Nakamura, 2009: Drought in the Southeastern United States: Causes, Variability over the Last Millennium, and the Potential for Future Hydroclimate Change. Journal of Climate, 22, 5021-5045. Secretary of the Interior, 2010: Addressing the Impacts of Climate Change on America s Water, Land, and Other Natural and Cultural Resources Washington, D.C., . Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Seiler, T., D. Staib, and M. Puttaiah, 2013: World Insurance in 2012: Progressing on the long and winding road to recovery. In: Sigma no. 3. Swiss Re, Zurich. Sharp, J.H., 2010: Estuarine oxygen dynamics: What can we learn about hypoxia from long-time records in the Delaware Estuary? Limnology and Oceanography, 55(2), 535-548. Sheffield, J., E.F. Wood, and M.L. Roderick, 2012: Little change in global drought over the past 60 years. Nature, 491(7424), 435-438. Shepherd, A., K.M. Gill, and S.B. Rood, 2010: Climate change and future flows of Rocky Mountain rivers: Converging forecasts from empirical trend projection and down-scaled global circulation modelling. Hydrological Processes, 24(26), 3864-3877. Shrestha, R.R., M.A. Schnorbus, A.T. Werner, and A.J. Berland, 2012: Modelling spatial and temporal variability of hydrologic impacts of climate change in the Fraser river basin, British Columbia, Canada. Hydrological Processes, 26(12), 1841-1861. Sillmann, J., V.V. Kharin, F.W. Zwiers, X. Zhang, and D. Bronaugh, 2013: Climate extremes indices in the CMIP5 multimodel ensemble: Part 2. Future climate projections. Journal of Geophysical Research: Atmospheres, 118, 2473-2493. Simmons, K.M. and D. Sutter, 2007: Tornado shelters and the manufactured home parks market. Natural Hazards, 43(3), 365-378. Simmons, T.A. (ed.), 1974: The Damnation of a Dam: The High Ross Dam Controversy. Diss. Master of Arts, Simon Fraser University, Department of Geography, British Columbia, Canada, . Sinervo, B., F. Méndez-de-la-Cruz, D.B. Miles, B. Heulin, E. Bastiaans, M. Villagrán-Santa Cruz, R. Lara- Resendiz, N. Martínez-Méndez, M.L. Calderón-Espinosa, R.N. Meza-Lázaro, H. Gadsden, L.J. Avila, M. Morando, I.J. De la Riva, P.V. Sepulveda, C.F.D. Rocha, N. Ibargüengoytía, C.A. Puntriano, M. Massot, V. Lepetz, T.A. Oksanen, D.G. Chapple, A.M. Bauer, W.R. Branch, J. Clobert, and J.W. Sites, 2010: Erosion of Lizard Diversity by Climate Change and Altered Thermal Niches. Science, 328(5980), 894-899. Singer, B.D., L.H. Ziska, D.A. Frenz, D.E. Gebhard, and J.G. Straka, 2005: Research note : Increasing Amb a 1 content in common ragweed (Ambrosia artemisiifolia) pollen as a function of rising atmospheric CO2 concentration. Functional Plant Biology, 32(7), 667-670. Skinner, W.R., E. Shabbar, M.D. Flannigan, and K. Logan, 2006: Large forest fires in Canada and the relationship to global sea surface temperatures. Journal of Geophysical Research, 111(D14106). Skoufias, E., K. Vinha, and H.V. Conroy, 2011: The Impacts of Climate Variability on Welfare in Rural Mexico. In: Policy Research Working Paper 5555. The World Bank, Poverty Reduction and Economic Management Network, Poverty Reduction and Equity Unit. Smith, J.A., M.L. Baeck, G. Villarini, and W.F. Krajewski, 2010: The hydrology and hydrometeorology of flooding in the Delaware River Basin. Journal of Hydrometeorology, 11(4), 841-859. Smith, P. and P.J. Gregory, 2013: Climate change and sustainable food production. Proceedings of the Nutrition Society, 1(1), 1-8. Smith, J.A., G. Villarini, and M.L. Baeck, 2011: Mixture Distributions and the Hydroclimatology of Extreme Rainfall and Flooding in the Eastern United States. Journal of Hydrometeorology, 12, 294-309. Smith, J.B., J.M. Vogel, and J.E.I. Cromwell, 2009: An Architecture for Government Action on Adaptation to Climate Change. An Editorial Comment. Climatic Change, 95(1-2), 53-61. Smolka, M.O. and A.A. Larangeira, 2008: Informality and poverty in Latin American urban policies. In: The New Global Frontier: Urbanization, Poverty and Environment in the 21st Century. Earthscan, London, UK, pp. 99- 114. Sobel, A.H., M. Biasutti, S.J. Camargo, and T.T. Creyts, 2010: Projected Changes in the Physical Climate of the Gulf Coast and Caribbean. Climatic Change, 112(3-4), 819-845. Sohngen, B. and R. Sedjo, 2005: Impacts of climate change on forest product markets: Implications for North American producers. The Forestry Chronicle, 81(5), 669-674. Solecki, W., 2012: Urban environmental challenges and climate change action in New York City. Environment and Urbanization, 24(2), 557-573. Sosa-Rodriguez, F.S., 2010: Impacts of Water-management Decisions on the Survival of a City: From Ancient Tenochtitlan to Modern Mexico City. Water Resources Development, 26(4), 675-687. Sosa-Rodriguez, F.S., 2013: From Federal to City Mitigation and Adaptation Strategies: Climate Change Policy in Mexico. Mitigation and Adaptation Strategies for Global Change, , 1-28. Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Soto-Pinto, L. and M. Anzueto, 2010: Carbon sequestration through agroforestry in indigenous communities of Chiapas, Mexico. Agroforestry Systems, 78(1), 39-51. Statistics Canada, 2012: Table 051- 0001 Estimates of population, by age group and sex for July 1, Canada, provinces and territories, annual Statistics Canada, Canada. Steiner, A.L., S. Tonse, R.C. Cohen, A.H. Goldstein, and R.a. Harley, 2006: Influence of future climate and emissions on regional air quality in California. Journal of Geophysical Research, 111, 1-22. Stewart, S.I., V.C. Radeloff, R.B. Hammer, and T.J. Hawbaker, 2006: Defining the Wildland-Urban Interface. Journal of Forestry, 105, 201-207. Stöckle, C.O., R.L. Nelson, S. Higgins, J. Brunner, G. Grove, R. Boydston, M. Whiting, and C. Kruger, 2010: Assessment of climate change impact on Eastern Washington agriculture. Climatic Change, 102(1-2), 77-102. Stratos Inc, 2011: Climate Change and Acid Rock Drainage - Risks for the Canadian Mining Sector. In: Mine Environment Neutral Drainage (MEND) Program Report 1.67.7. MEND. Suddick, E.C., K.M. Scow, W.R. Horwath, L.E. Jackson, D.R. Smart, J. Mitchell, and J. Six, 2010: The potential for California agricultural crop soils to reduce greenhouse gas emissions: A holistic evaluation. Advances in Agronomy, 107, 124-162. Sun, G., S.G. McNulty, J.A. Moore Myers, and E.C. Cohen, 2008: Impacts of multiple stresses on water demand and supply across the southeastern United States. Journal of the American Water Resources Association, 44(6), 1441-1457. Swanson, D., J. Hiley, and H.D. Venema, 2007: Indicators of Adaptive Capacity to Climate Change for Agriculture in the Prairie Region of Canada: An analysis based on Statistics Canada s Census of Agriculture. In: Working Paper for the Prairie Climate Resilience Project. International Institute for Sustainable Development, pp. 63. Sydneysmith, R., M. Andrachuk, B. Smit, and G.K. Hovelsrud, 2010: Vulnerability and Adaptive Capacity in Arctic Communities. In: Adaptive Capacity and Environmental Governance. [Armitage, D. and R. Plummer(eds.)]. Springer-Verlag, Berlin, pp. 133-156. Tabachnick, W.J., 2010: Challenges in predicting climate and environmental effects on vector-borne disease episystems in a changing world. The Journal of Experimental Biology, 213(6), 946-54. Tagaris, E., K. Liao, K. Manomaiphiboon, S. He, J. Woo, P. Amar, and A.G. Russell, 2008: The role of climate and emission changes in future air quality over southern Canada and northern Mexico. Atmospheric Chemistry and Physics, 8(14), 3973-3983. Tagaris, E., K. Liao, A.J. Delucia, L. Deck, P. Amar, and A.G. Russell, 2009: Potential impact of climate change on air pollution-related human health effects. Environmental Science & Technology, 43(13), 4979-88. Tagaris, E., K. Liao, A.J. DeLucia, L. Deck, P. Amar, and A.G. Russell, 2010: Sensitivity of air pollution-induced premature mortality to precursor emissions under the influence of climate change. International Journal of Environmental Research and Public Health, 7(5), 2222-37. Tai, A.P.K., L.J. Mickley, and D.J. Jacob, 2010: Correlations between fine particulate matter (PM2.5) and meteorological variables in the United States: Implications for the sensitivity of PM2.5 to climate change. Atmospheric Environment, 44(32), 3976-3984. Tambo, J.A. and T. Abdoulaye, 2012: Climate change and agricultural technology adoption: the case of drought tolerant maize in rural Nigeria. Mitigation and Adaptation Strategies for Global Change, 17(3), 277-292. Taner, M.T., J.N. Carleton, and M. Wellman, 2011: Integrated model projections of climate change impacts on a North American lake. Ecological Modelling, 222(18), 3380-3393. Tanzeeba, S. and T.Y. Gan, 2012: Potential impact of climate change on the water availability of South Saskatchewan River Basin. Climatic Change, 112(2), 355-386. Tao, Z., A. Williams, H. Huang, M. Caughey, and X. Liang, 2007: Sensitivity of U.S. surface ozone to future emissions and climate changes. Geophysical Research Letters, 34(8), 1-5. Tarnoczi, T., 2011: Transformative learning and adaptation to climate change in the Canadian Prairies agro- ecosystem. Mitigation and Adaptation Strategies for Global Change, 16(4), 387-406. Tarnoczi, T.J. and F. Berkes, 2010: Sources of information for farmers adaptation practices in Canada s Prairie agro-ecosystem. Climatic Change, 98(1), 299-305. Tebaldi, C., B.H. Strauss, and C.E. Zervas, 2012: Modelling sea level rise impacts on storm surges along US coasts. Environmental Research Letters, 7(1), 014032. Teixeira, E.I., G. Fischer, H. van Velthuizen, C. Walter, and F. Ewert, 2013: Global hot-spots of heat stress on agricultural crops due to climate change. Agricultural and Forest Meteorology, 170, 206-215. Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Teranishi, H., T. Katoh, K. Kenda, and S. Hayashi, 2006: Global warming and the earlier start of the Japanese-cedar (Cryptomeria japonica) pollen season in Toyama, Japan. Aerobiologia, 22(2), 90-94. Terrier, A., M.P. Girardin, C. Périé, P. Legendre, and Y. Bergeron, 2013: Potential changes in forest composition could reduce impacts of climate change on boreal wildfires. Ecological Applications, 23(1), 21-35. The White House, 2009: Federal Leadership in Environmental, Energy, and Economic Performance - Executive Order 13514 Executive Office of the President, Washington, D.C., . Theobald, D.M. and W.H. Romme, 2007: Expansion of the US wildland-urban interface. Landscape and Urban Planning, 83, 340-354. Thomas, M.K., D.F. Charron, D. Waltner-Toews, C.J. Schuster, A.R. Maarouf, and J.D. Holt, 2006: A role of high impact weathe revents in water borne disease outbreaks in Canada. International Journal of Environmental Health Research, 16(3), 167-180. Tingley, M.W., W.B. Monahan, S.R. Beissinger, and C. Moritz, 2009: Birds track their Grinnellian niche through a century of climate change. Proceedings of the National Academy of Sciences, 106(Supplement 2), 19637- 19643. Tong, S.T.Y., Y. Sun, T. Ranatunga, J. He, and Y.J. Yang, 2012: Predicting plausible impacts of sets of climate and land use change scenarios on water resources. Applied Geography, 32(2), 477-489. Trainor, S.F., M. Calef, D. Natcher, F.S. Chapin III, A.D. McGuire, O. Huntington, P. Duffy, T.S. Rupp, L. DeWilde, M. Kwart, N. Fresco, and A.L. Lovecraft, 2009: Vulnerability and adaptation to climate-related fire impacts in rural and urban interior Alaska. Polar Research, 28, 100-118. Transportation Research Board, 2008: Potential Impacts of Climate Change on U.S. Transportation. National Research Council of the National Academy of Sciences, Washington, D.C. Trapp, R.J., N.S. Diffenbaugh, and A. Gluhovsky, 2009: Transient response of severe thunderstorm forcing to elevated greenhouse gas concentrations. Geophysical Research Letters, 36(1), L01703. Trejo, I., E. Martínez-Meyer, E. Calixto-Pérez, S. Sánchez-Colón, R. Vázquez de La Torre, and L. Villers-Ruiz, 2011: Analysis of the effects of climate change on plant communities and mammals in México. Atmosfera, 24(1), 1-14. Tremblay, L., M. Larocque, F. Anctil, and C. Rivard, 2011: Teleconnections and Interannual Variability in Canadian Groundwater Levels. Journal of Hydrology, 410, 178-188. Trumpickas, J., B.J. Shuter, and C.K. and Minns, 2009: Forecasting impacts of climate change on Great Lakes surface water temperatures. Journal of Great Lakes Research, 35(3), 454-463. Tu, J., 2009: Combined impact of climate and land use changes on streamflow and water quality in eastern Massachusetts, USA. Journal of Hydrology, 379(3-4), 268-283. Tufts, S., 2010: Tourism, climate change and the missing worker: uneven impacts, institutions and response. In: What Do We Know? What Do We Need to Know? The State of Canadian Research on Work, Employment and Climate Change. [Lipsig-Mummé, C. (ed.)]. Work in a Warming World Research Programme, York University, Toronto, pp. 80-108. U.S. Army Corps of Engineers and Bonneville Power Administration, 2013: Columbia River Treaty 2014/2024 Review U.S. Army Corps of Engineers; Bonneville Power Administration, . U.S. Census Bureau, 2011: Statistical Abstract of the United States: 2012. Washington, D.C. U.S. Department of Agriculture, Economic Research Service, 2012: U.S. Drought 2012: Farm and Food Impacts U.S. Department of Agriculture, Economic Research Service, Washington, D.C., . U.S. Department of Energy, 2013: U.S. Energy Sector Vulnerabilities to Climate Change and Extreme Weather. U.S. Department of Energy, Washington, D.C. U.S. Environmental Protection Agency, 2004: National Water Quality Inventory: Report to Congress for the 2004 Reporting Cycle
. U.S. Environmental Protection Agency, Washington, D.C., pp. 1-37. U.S. Environmental Protection Agency, 2008: Review of the Impacts of Climate Variability and Change on Aeroallergens and Their Associated Effects. U.S. Environmental Protection Agency, Washington, D.C., pp. 1- 125. U.S. Environmental Protection Agency and Secretaría de Medio Ambiente y Recursos Naturales, 2011: State of the Border Region Indicators Report 2010. Secretaría de Medio Ambiente y Recursos Naturales, Tlalpan, Mexico, pp. 1-99. U.S. Environmental Protection Agency, 2012: U.S.-Mexico Environmental Program: Border 2012: A Mid-Course Refinement (2008-2012). U.S. Environmental Protection Agency, Washington, D.C. Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 U.S. Environmental Protection Agency, National Water Program, 2011: U.S. EPA National Water Program Strategy: Response to Climate Change 2010 - 2011 National and Regional Highlights of Progress. U.S. Environmental Protection Agency, Washington, D.C. U.S. Environmental Protection Agency, Office of Atmospheric Programs, 2012: Climate Change Indicators in the United States, 2012 . U.S Environmental Protection Agency, Washington, D.C., pp. 72-73. U.S. Forest Service, 2010: National Roadmap for Responding to Climate Change. U.S. Department of Agriculture, Forest Service, Washington, D.C., pp. 30. U.S. Global Change Research Program, 2009: Global Climate Change Impacts in the United States. Cambridge University Press, New York, USA, pp. 1-192. U.S. Government Accountability Office, 2007: Port Risk Management: Additional Federal Guidance Would Aid Ports in Disaster Planning and Recovery. U.S. Government Accountability Office, Washington, D.C., pp. 1-57. U.S. Government Accountability Office, 2009: Climate Change Adaptation: Strategic Federal Planning could help government officials make more informed decisions. In: Report to the Chairman, Select Committee on Energy Independence and Global Warming, House of Representatives [Government Accounting Office (ed.)]. Congressional Report. Government of United States, Washington, D.C., pp. 1-86. Urban, D., M.J. Roberts, W. Schlenker, and D.B. Lobell, 2012: Projected temperature changes indicate significant increase in interannual variability of U.S. maize yields. Climatic Change, 112, 525-533. Ureta, C., E. Martínez Meyer, H.R. Perales, and E.R. Álvarez Buylla, 2012: Projecting the effects of climate change on the distribution of maize races and their wild relatives in Mexico. Global Change Biology, 18(3), 1073-1082. van Mantgem, P.J., N.L. Stephenson, J.C. Byrne, L.D. Daniels, J.F. Franklin, P.Z. Fulé, M.E. Harmon, A.J. Larson, J.M. Smith, A.H. Taylor, and T.T. Veblen, 2009: Widespread Increase of Tree Mortality Rates in the Western United States. Science, 323(5913), 521-524. Vano, J., N. Voisin, L. Cuo, A. Hamlet, M. Elsner, R. Palmer, A. Polebitski, and D. Lettenmaier, 2010a: Climate change impacts on water management in the Puget Sound region, Washington State, USA. Climatic Change, 102(1-2), 261-286. Vano, J.A., M.J. Scott, N. Voisin, C.O. Stöckle, A.F. Hamlet, K. Mickelson, and D.P. Lettenmaier, 2010b: Climate change impacts on water management and irrigated agriculture in the Yakima River Basin, Washington, USA. Climatic Change, 102(1-2), 287-317. Varady, R. and E. Ward, 2009: Transboundary conservation in context: What drives environmental change? Conservation of Shared Environments: Learning from the United States and Mexico, , 9-22. Vasseur, L. and N. Catto, 2008: Atlantic Canada. In: From Impacts to Adaptation: Canada in a Changing Climate 2007. [Lemmen, D.S., F.J. Warren, J. Lacroix, and E. Bush(eds.)]. Government of Canada, Ottawa, ON, pp. 119-170. Vasseur, L., 2011: Moving from Research into Action on Issues of ClimateChange for a Canadian Community: Integration ofSciences into Decision Making. The International Journal of Climate Change: Impacts & Responses, 2(4), 116-126. Veron, J.E.N., O. Hoegh-Guldberg, T.M. Lenton, J.M. Lough, D.O. Obura, P. Pearce-Kelly, C.R.C. Sheppard, M. Spalding, M.G. Stafford-Smith, and A.D. Rogers, 2009: The coral reef crisis: The critical importance of <350 ppm CO2. Marine Pollution Bulletin, 58(10), 1428-1436. Vespa, J., J. Lewis, and R. Kreider, 2013: America's Families and Living Arrangements: 2012 U.S. Census Bureau, Washington, D.C., . Villarini, G., J.A. Smith, F. Serinaldi, J. Bales, P.D. Bates, and W.F. Krajewski, 2009: Flood frequency analysis for nonstationary annual peak records in an urban drainage basin. Advances in Water Resources, 32(8), 1255-1266. Villarini, G. and J.A. Smith, 2010: Flood peak distributions for the eastern United States. Water Resources Research, 46(6). Villarini, G., J.A. Smith, M.L. Baeck, and W.F. Krajewski, 2011: Examining Flood Frequency Distributions in the Midwest U.S. Journal of the American Water Resources Association, 47(3), 447-463. Villeneuve, P.J. and R.T. Burnett, 2003: A time-series study of air pollution, socioeconomic status, and mortality in Vancouver, Canada. Journal of Exposure Science and Environmental Epidemiology, 13(6), 427-435. Villeneuve, P.J., M. Doiron, D. Stieb, R. Dales, R.T. Burnett, and R. Dugandzic, 2006: Is outdoor air pollution associated with physician visits for allergic rhinitis among the elderly in Toronto, Canada? Allergy, 61(6), 750- 758. Villers-Ruiz, L. and D. Castaneda-Aguado, In press: Species and Plant Communities Reorganization in the Trans- Mexican Volcanic Belt under Climate Change Conditions . Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Villers-Ruíz, L. and J. Hernández-Lozano, 2007: Incendios forestales y el fenómeno de El Nino en México.Proceedings of IV Conferencia Internacional sobre Incendios Forestales, May 13-17, 2007, Sevilla, Espana, pp. 1-10. von Peter, G., S. von Dahlen, and S. Saxena, 2012: Unmitigated Disasters? New evidence on the macroeconomic cost of natural catastrophes. In: BIS Working Papers No 394. Bank for International Settlements. Vose, J.M., D.L. Peterson, and T. Patel-Weynand, 2012: Effects of climatic variability and change on forest ecosystems: a comprehensive science synthesis for the U.S. In: General Technical Report PNW-GTR-870. U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station, Portland, OR, pp. 1-265. Vrolijks, L., A. Spatafpre, and A.S. Mittel, 2011: Comparative research on the adaptation strategies of ten urban climate plans. In: Resilient Cities: Cities and Adaptation to Climate Change [Otto-Zimmerman, K. (ed.)]. Proceedings of Proceedings of the Global Forum 2010, Local Sustainability, pp. 193-203. Walsh, C.J., T.D. Fletcher, and A.R. Ladson, 2005: Stream restoration in urban catchments through redesigning stormwater systems: looking to the catchment to save the stream. Journal of the North American Benthological Society, 24(3), 690-705. Walthall, C.L., J. Hatfield, P. Backlund, L. Lengnick, E. Marshall, M. Walsh, S. Adkins, M. Aillery, E.A. Ainsworth, C. Ammann, C.J. Anderson, I. Bartomeus, L.H. Baumgard, F. Booker, B. Bradley, D.M. Blumenthal, J. Bunce, K. Burkey, S.M. Dabney, J.A. Delgado, J. Dukes, A. Funk, K. Garrett, M. Glenn, D.A. Grantz, D. Goodrich, S. Hu, R.C. Izaurralde, R.A.C. Jones, S.-. Kim, A.D.B. Leaky, K. Lewers, T.L. Mader, A. McClung, J. Morgan, D.J. Muth, M. Nearing, D.M. Oosterhuis, D. Ort, C. Parmesan, W.T. Pettigrew, W. Polley, R. Rader, C. Rice, M. Rivington, E. Rosskopf, W.A. Salas, L.E. Sollenberger, R. Srygley, C. Stöckle, E.S. Takle, D. Timlin, J.W. White, R. Winfree, L. Wright-Morton, and L.H. Ziska, 2012: Climate Change and Agriculture in the United States: Effects and Adaptation. In: USDA Technical Bulletin 1935, Washington, D.C., pp. 1-186. Wang, X.L., V.R. Swail, and F.W. Zwiers, 2006: Climatology and changes of extratropical cyclone activity: Comparison of ERA-40 with NCEP-NCAR reanalysis for 1958-2001. Journal of Climate, 19, 3145-3166. Waring, K.M., D.M. Reboletti, L.A. Mork, C. Huang, R.W. Hofstetter, A.M. Garcia, P.Z. Fulé, and T.S. Davis, 2009: Modeling the impacts of two bark beetle species under a warming climate in the southwestern USA: Ecological and economic consequences. Environmental Management, 44, 824-835. Warren, F.J. and P. Egginton, 2008: Chapter 2: Background Information - Concepts, Overviews and Approaches. In: From Impacts to Adaptation: Canada in a Changing Climate 2007. Government of Canada, Ottawa, pp. 27-56. Washington State Built Environment: Infrastructure & Communities Topic Advisory Group, 2011: Interim Recommendations of the Built Environment: Infrastructure & Communities Topic Advisory Group. In: Washington State Climate Change Response Strategy. Department of Ecology, State of Washington, Olympia, WA, pp. 1-53. Washington State Human Health and Security Topic Advisory Group, 2011: Topic Advisory Group (TAG) Report- TAG 2 Human Health and Security. In: Washington State Climate Change Response Strategy. Department of Ecology, State of Washington, Olympia, WA, pp. 1-17. Washington State Natural Resources Working Lands and Waters Topic Advisory Group, 2011: Interim Recommendations of the Natural Resources: Working Lands and Waters Topic Advisory Group. In: Washington State Climate Change Response Strategy. Department of Ecology, State of Washington, Olympia, WA, pp. 1- 44. Washington State Species, Habitats and Ecosystems Topic Advisory Group, 2011: Interim Recommendations from Topic Advisory Group 3 Species, Habitats and Ecosystems (TAG3). In: Washington State Integrated Climate Change Response Strategy. Department of Ecology, State of Washington, Olympia, WA, pp. 1-95. Wayne, P., S. Foster, J. Connolly, F. Bazzaz, and P. Epstein, 2002: Production of allergenic pollen by ragweed (Ambrosia artemisiifolia L.) is increased in CO2-enriched atmospheres. Annals of Allergy, Asthma & Immunology, 88(3), 279-82. Wehbe, M., H. Eakin, R. Seiler, M. Vinocur, C. Avila, and C. Marutto, 2008: Local perspectives on adaptation to climate change: lessons from Mexico and Argentina. In: Climate Change and Adaptation. Earthscan, London, UK, pp. 296-314. Wehner, M., D.R. Easterling, J.H. Lawrimore, R.R. Heim, R.S. Vose, and B.D. Santer, 2011: Projections of future drought in the continental United States and Mexico. Journal of Hydrometeorology, 12(6), 1359-1377. Weisler, R.H., J.G. Barbee, and M.H. Townsend, 2006: Mental health and recovery in the Gulf Coast after Hurricanes Katrina and Rita. The Journal of the American Medical Association, 296(5), 585-8. Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Weiss, J.L., J.T. Overpeck, and B. Strauss, 2011: Implications of Recent Sea Level Rise Science for Low-elevation Areas in Coastal Cities of the Conterminous U.S.A. Climatic Change, 105, 635-645. Westerling, A.L., H.G. Hidalgo, D.R. Cayan, and T.W. Swetnam, 2006: Warming and Earlier Spring Increase Western U.S. Forest Wildfire Activity. Science, 313, 940-943. Wheaton, E., G. Koshida, B. Bonsal, T. Johnston, W. Richards, and V. Wittrock, 2007: Agricultural Adaptation to Drought (ADA) in Canada: The Case of 2001 and 2002. In: SRC Publication No. 11927-1E07. Saskatchewan Research Council, Saskatoon, SK. Wheaton, E., S. Kulshreshtha, W. Eilers, and V. Wittrock, 2010: Environmental sustainability of agriculture in the context of adapting to a changing climate.Proceedings of OECDINEA-FAO Workshop, Rome, Italy, . Whitener, L.A. and T. Parker, 2007: Policy Options for a Changing Rural America. Perspectives on Food and Farm Policy: Rural Development & Energy, 5, 58-65. White-Newsome, J.L., B.N. Sánchez, O. Jolliet, Z. Zhang, E.A. Parker, J. Timothy Dvonch, and M.S. O'Neill, 2012: Climate change and health: Indoor heat exposure in vulnerable populations. Environmental Research, 112, 20- 27. Wiedinmyer, C. and M.D. Hurteau, 2010: Prescribed fire as a means of reducing forest carbon emissions in the Western United States. Environment, Science and Technology, 44, 1926-1932. Wilbanks, T., D. Bilello, D. Schmalzer, D. Arent, J. Buizer, H. Chum, J. Dell, J. Edmonds, G. Franco, R. Jones, S. Rose, N. . Roy, A. Sanstad, S. Seidel, J. Weyant, and D. Wuebbles, 2012: Climate Change and Energy Supply and Use: Technical Report for the U.S. Department of Energy in Support of the National Climate Assessment. Oak Ridge National Laboratory, Oak Ridge, Tennessee. Wilbanks, T.J., V. Bhatt, D.E. Bilello, S.R. Bull, J. Ekmann, W.C. Horak, Y.J. Huang, M.D. Levine, M.J. Sale, D.K. Schmalzer, and M.J. Scott, 2008: Effects of Climate Change on Energy Production and Use in the United States. Synthesis and Assessment Product 4.5
. In: Report by the U.S. Climate Change Science Program and the Subcommittee on Global Change Research. Department of Energy, Office of Biological & Environmental Research, Washington, D.C. Wilder, M., C.A. Scott, N.P. Pablos, R.G. Varady, G.M. Garfin, and J. McEvoy, 2010: Adapting Across Boundaries: Climate Change, Social Learning, and Resilience in the US Mexico Border Region. Annals of the Association of American Geographers, 100(4), 917-928. Wilder, M., G. Garfin, P. Ganster, H. Eakin, P. Romero-Lankao, F. Lara-Valencia, A.A. Cortez-Lara, S. Mumme, C. Neri, and F. Munoz-Arriola, 2013: Chapter 16: Climate Change and U.S.-Mexico Border Communities. In: Assessment of Climate Change in the Southwest United States: a Report Prepared for the National Climate Assessment. [Garfin, G., A. Jardine, R. Merideth, M. Black, and S. LeRoy(eds.)]. Island Press, Washington, D.C., . Williams, A.P., C.D. Allen, C.I. Millar, T.W. Swetnam, J. Michaelson, C.J. Still, and S.W. Leavitt, 2010: Forest responses to increasing aridity and warmth in the southwestern United States. Proceedings of the National Academy of Sciences, 107(50), 21289-21294. Williamson, T.B., M. Patriquin, and V. Wittrock, 2008: Chapter 5: Adaptive capacity deficits of human populations in the Canadian boreal plains ecozone: assessment and issues. In: Limited Report: Climate Change Adaptive Capacity of Forestry Stakeholders in the Boreal Plains Ecozone. Climate Change Impacts and Adaptation Program Project A1383 [Johnston, M., Williamson, T., Wheaton, E., Wittrock, V., Nelson, H., Hesseln, H., Vandamme, L., Pittman, J. and Lebel, M. (eds.)]. Saskatchewan Research Council, Saskatoon, SK, pp. 46-81. Williamson, T.B., S.J. Colombo, P.N. Duinker, P.A. Gray, R.J. Hennessey, D. Houle, M.H. Johnston, A.E. Ogden, and D.L. Spittlehouse, 2009: Climate change and Canada s forests: from impacts to adaptation. Edmonton, Canada, Sustainable Forest Management Network and Natural Resources Canada, Canadian Forest Service, Northern Forestry Centre ed., pp. 104. Wilson, C.O. and Q. Weng, 2011: Simulating the impacts of future land use and climate changes on surface water quality in the Des Plaines River watershed, Chicago Metropolitan Statistical Area, Illinois. Science of the Total Environment, 409(20), 4387-4405. Wilson, K., 2009: Climate change and the spread of infectious ideas. Ecology, 90(4), 901-2. Winder, R., E.A. Nelson, and T. Beardmore, 2011: Ecological implications for assisted migration in Canadian forests. The Forestry Chronicle, 87(6), 731-744. Wittrock, V., S.N. Kulshreshtha, and E. Wheaton, 2011: Canadian prairie rural communities: their vulnerabilities and adaptive capacities to drought. Mitigation and Adaptation Strategies for Global Change, 16, 267-290. Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wolf, J., W.N. Adger, I. Lorenzoni, V. Abrahamson, and R. Raine, 2010: Social capital, individual responses to heat waves and climate change adaptation: An empirical study of two U.K. cities. Global Environmental Change, 20, 44-52. Wolfe, D.W., L. Ziska, C. Petzoldt, A. Seaman, L. Chase, and K. Hayhoe, 2008: Projected change in climate thresholds in the Northeastern U.S.: implications for crops, pests, livestock, and farmers. Mitigation and Adaptation Strategies for Global Change, 13, 555-575. Woodhouse, C., D.M. Meko, G.M. MacDonald, D.W. Stahle, and E.R. Cook, 2010: 1,200-year perspective of 21st century drought in southwestern North America. Proceedings of the National Academies of Science, 107, 21283-21288. Wootton, J.T., C.A. Pfister, and J.D. Forester, 2008: Dynamic patterns and ecological impacts of declining ocean pH in a high resolution multi-year dataset. Proceedings of the National Academy of Sciences, 105, 18848-18853. World Bank, 2009: Globalisation, Wages, and the Quality of Jobs Five Country Studies. [Robertson, R., Brown, D., Pierre, G. and Sanchez-Puerta, L. (eds.)]. The World Bank, Washington, D.C. World Bank, 2010: Natural Hazards, UnNatural Disasters: The Economics of Effective Prevention. The International Bank for Reconstruction and Development, Washington, D.C. World Bank, 2011: Global Economic Prospects: Maintaining Progress Amid Turmoil Volume 3. The International Bank for Reconstruction and Development / The World Bank, Washington, D.C. Wright, L., P. Chinowsky, K. Strzepek, R. Jones, R. Streeter, J.B. Smith, J.-. Mayotte, A. Powell, L. Jantarasami, and W. Perkins, 2012: Estimated effects of climate change on flood vulnerability of U.S. Bridges. Mitigation and Adaptation Strategies for Global Change, 17(8), 939-955. Wu, S., 2010: Potential impact of climate change on flooding in the upper great miami river watershed, ohio, USA: A simulation-based approach. [Impact potentiel du changement climatique sur les inondations dans le bassin supérieur de la Riviere Great Miami, Ohio, Etats-Unis: Une approche basée sur la simulation]. Hydrological Sciences Journal, 55(8), 1251-1263. Wu, S., L.J. Mickley, E.M. Leibensperger, D.J. Jacob, D. Rind, and D.G. Streets, 2008: Effects of 2000â 2050 global change on ozone air quality in the United States. Journal of Geophysical Research, 113, 1-12. Yanchuk, A. and G. Allard, 2009: Tree improvement programmes for forest health - can they keep pace with climate changes? Unasylva, 60(231-232), 50-56. Yang, L., J.A. Smith, D.B. Wright, M.L. Baeck, G. Villarini, F. Tian, and H. Hu, 2013: Urbanization and Climate Change: An Examination of Nonstationarities in Urban Flooding. Journal of Hydrometeorology, (2013). Yin, J., M.E. Schlesinger, and R.J. Stouffer, 2009: Model projectsion of rapid sea-level rise on the northeast coast of the United States. Nature Geoscience, 2(4), 262-266. Yin, J.H., 2005: A consistent poleward shift of the storm tracks in simulations of 21st century climate. Geophysical Research Letters, 32(18), L18701. Zhang, X.C. and M.A. Nearing, 2005: Impact of climate change on soil erosion, runoff, and wheat productivity in central Oklahoma. Catena, 61, 185-195. Ziska, L., K. Knowlton, C. Rogers, D. Dalan, N. Tierney, M. Ann, W. Filley, J. Shropshire, L.B. Ford, C. Hedberg, P. Fleetwood, K.T. Hovanky, and T. Kavanaugh, 2011: Recent warming by latitude associated with increased length of ragweed pollen season in central North America. Proceedings of the National Academies of Science, 108(10), 4248-4251. Ziska, L.H. and F.A. Caulfield, 2000: Rising CO2 and pollen production of common ragweed (Ambrosia artemisiifolia), a known allergy-inducing species: implications for public health. Australian Journal of Plant Physiology, 27, 893-898. Ziska, L.H., D.E. Gebhard, D.a. Frenz, S. Faulkner, B.D. Singer, and J.G. Straka, 2003: Cities as harbingers of climate change: Common ragweed, urbanization, and public health. Journal of Allergy and Clinical Immunology, 111(2), 290-295. Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 26-1: Key risks from climate change and the potential for risk reduction through adaptation. Key risks are identified based on assessment of the literature and expert judgments made by authors of this chapter, with supporting evaluation of evidence and agreement in the referenced chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels are presented for the near-term era of committed climate change (here, for 2030-2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the longer-term era of climate options (here, for 2080-2100), for global mean temperature increase of 2°C and 4°C above preindustrial levels. For each timeframe, risk levels are estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols. Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 26-1 Detection and attribution of climate change impacts. Comparisons of the adequacy of currently available data to detect trends and the degree of understanding of causes of those changes in climatic extreme events in the United States (left; Peterson et al., 2013) and degree of understanding of the climate influence in key impacts in North America (right). Note that climate influence means that the impact has been documented to be sensitive to climate, not that it has been attributed to climate change. Filled boxes indicate that formal detection and attribution to climate change has been performed for the given impact; shaded boxes indicate that a trend has been detected from background variability in the given impact, but formal attribution to climate change has not occurred and the trend could be due to other drivers; and open boxes indicate that a trend has not currently been detected. Key impacts are: 1) earlier peak flow of snowmelt run-off in snow-dominated streams and rivers in western North America [26.3.1], 2) declines in the amount of water stored in spring snowpack in snow-dominated areas of western North America [26.3.1], 3) northward and upward shifts in species distributions in multiple taxa of terrestrial species, although not all taxa and regions [26.4.1], 4) increases in coastal flooding [26.8.1], 5) increases in wildfire activity, including fire season length and area burned by wildfires in the western United States and boreal Canada [Box 26-2], 6) storm-related disaster losses in the United States (most of the increase in insurance claims paid has been attributed to increasing exposure of people and assets in areas of risk) [26.7.6.1, 26.8.1], 7) increases in bark beetle infestation levels in pine tree species in western North America [26.4.2.1], 8) yield increases due in part to increasing temperatures in Canada and higher precipitation in the US; yield variances attributed to climate variability in Ontario and Quebec; yield losses attributed to climate-related extremes across North America [26.5.1], 9) changes in storm-related mortality in the United States [26.6.1.2], 10) changes in heat-related mortality in the United States [26.6.1.2], 11) increases in tree mortality rates in old-growth forests in the western United States and western Canada from 1960-2007 [26.4.2.1], 12) changes in flooding in some urban areas due to extreme rainfall [26.3.1, 26.8.2.1], 13) increase in water supply shortages due to drought [26.8.1, 26.3], and 14) changes in cold- related heat mortality [26.6.1.2]. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 85 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 26-2: Extreme events illustrating vulnerabilities for Mexico, the United States, and Canada. This figure offers a graphic illustration of location of extreme events and relevant vulnerability trends. The observed extreme events have not been attributed to anthropogenic climate change, yet they are climate-sensitive sources of impact illustrating vulnerability of exposed systems, particularly if projected future increases in the frequency and/or intensity of such events should materialize. The figure includes: a) A map (bottom) with population density at 1km resolution highlighting exposure and represented using 2011 Landscan data (Bright et al., 2012). b) A map (top) with significant weather events taking place during 1993-2012. The map only includes disasters with overall losses of more than $1 billion US dollars in US, or more than $500 million US dollars in Mexico and Canada, adjusted to 2012 values (Source: (NatCatSERVICE, 2010). Hence, it does not include the occurrence of disasters of small and medium impact, and it does not capture the impacts of disasters on populations livelihoods and wellbeing. Disasters represented by points that are located at the approximate geographic center of affected regions, frequently span more than one subnational jurisdiction (e.g., the 2012 drought affected 12 Mexican states, Annex Table). c) Four panels (right) with trends in socio-demographic indicators used in the literature to measure vulnerability to hazards (Romero-Lankao et al., 2012): poverty rates, percentage of elderly, GDP per capita and total population (Sources: Comisión Económica para América Latina y el Caribe; U.S. Census Bureau, 2011; Statistics Canada, 2012). Subject to Final Copyedit 86 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 26-3: Observed and projected Changes in annual temperature and precipitation. (Top panel, left) observed temperature trends from 1901-2012 determined by linear regression [WGI AR5 Figures SPM.1 and 2.21]. (Bottom panel, left) Observed precipitation change from 1951-2010 determined by linear regression. [WGI AR5 Figure SPM.2] For observed temperature and precipitation, trends have been calculated where sufficient data permits a robust estimate (i.e., only for grid boxes with greater than 70% complete records and more than 20% data availability in the first and last 10% of the time period). Other areas are white. Solid colors indicate areas where change is significant at the 10% level. Diagonal lines indicate areas where change is not significant (Top and bottom panel, right) CMIP5 multi-model mean projections of annual average temperature changes and average percent change in annual mean precipitation for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 26 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 variability, and >90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where >66% of models show change greater than the baseline variability and >66% of models agree on sign of change. Gray indicates areas with divergent changes, where >66% of models show change greater than the baseline variability, but <66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in >66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5. [Boxes 21-3 and CC-RC] Figure 26-4: Projected changes in extremes in North America. (a) The percentage of years in the 2046 2065 period of RCP8.5 in which the summer temperature is greater than the respective maximum summer temperature of the 1986 2005 baseline period (Diffenbaugh and Giorgi, 2012). (b) The percentage of years in the 2080-2099 period of RCP8.5 in which the summer precipitation is less than the respective minimum summer precipitation of the 1986- 2005 baseline period (Diffenbaugh and Giorgi, 2012) (c) The percentage difference in the 20-year return value of annual precipitation extremes between the 2046-2065 period of RCP4.5 and the 1986-2005 baseline period (from (Kharin et al., 2013). The hatching indicates areas where the differences are not significant at the 5% level. (d) The percentage of years in the 2070-2099 period of RCP8.5 in which the March snow water equivalent is less than the respective minimum March snow water equivalent of the 1976 2005 period (Diffenbaugh et al., 2012). The black (white) stippling indicate areas where the multimodel mean exceeds 1.0 (2.0) standard deviations of the multi-model spread. (a-d) The RCPs and time periods are those used in the peer-reviewed studies in which the panels appear. The 2046-2065 period of RCP8.5 and the 2046-2065 period of RCP4.5 exhibit global warming in the range of 2-3C above the pre-industrial baseline (WGI Fig. 12.40). The 2080-2099 and 2070-2099 periods of RCP8.5 exhibit global warming in the range of 4-5C above the pre-industrial baseline (WGI Fig. 12.40). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 27. Central and South America Coordinating Lead Authors Graciela Magrin (Argentina), José Marengo (Brazil) Lead Authors Jean-Phillipe Boulanger (France), Marcos S. Buckeridge (Brazil), Edwin Castellanos (Guatemala), Germán Poveda (Colombia), Fabio R. Scarano (Brazil), Sebastián Vicuna (Chile) Contributing Authors Eric Alfaro (Costa Rica), Fabien Anthelme (France), Jonathan Barton (UK), Nina Becker (Germany), Arnaud Bertrand (France), Ulisses Confalonieri (Brazil), Carlos Demiguel (Spain), Bernard Francou (France), Rene Garreaud (Chile), Inigo Losada (Spain), Melanie McField (USA), Carlos Nobre (Brazil), Patricia Romero Lankao (Mexico), Paulo Saldiva (Brazil), Jose Luis Samaniego (Mexico), Amanda Pereira de Souza (Brazil), María Travasso (Argentina), Ernesto Viglizzo (Argentina), Alicia Villamizar (Venezuela) Review Editors Leonidas Osvaldo Girardin (Argentina), Jean Pierre Ometto (Brazil) Volunteer Chapter Scientist Nina Becker (Germany) Contents Executive Summary 27.1. Introduction 27.1.1. The Central and South America Region 27.1.2. Summary of the AR4 and SREX Findings 27.1.2.1. AR4 Findings 27.1.2.2. SREX Findings 27.2. Major Recent Changes and Projections in the Region 27.2.1. Climatic Stressors 27.2.1.1. Climate Trends, Long-term Changes in Variability, and Extremes 27.2.1.2. Climate Projections 27.2.2. Non-Climatic Stressors 27.2.2.1. Trends and Projections in Land Use and Land Use Change 27.2.2.2. Trends and Projections in Socioeconomic Conditions 27.3. Impacts, Vulnerabilities and Adaptation Practices 27.3.1. Freshwater Resources 27.3.1.1. Observed and Projected Impacts and Vulnerabilities 27.3.1.2. Adaptation Practices 27.3.2. Terrestrial and Inland Water Systems 27.3.2.1. Observed and Projected Impacts and Vulnerabilities 27.3.2.2. Adaptation Practices 27.3.3. Coastal Systems and Low-Lying Areas 27.3.3.1. Observed and Projected Impacts and Vulnerabilities 27.3.3.2. Adaptation Practices 27.3.4. Food Production Systems and Food Security 27.3.4.1. Observed and Projected Impacts and Vulnerabilities 27.3.4.2. Adaptation Practices Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 27.3.5. Human Settlements, Industry, and Infrastructure 27.3.5.1. Observed and Projected Impacts and Vulnerabilities 27.3.5.2. Adaptation Practices 27.3.6. Renewable Energy 27.3.6.1. Observed and Projected Impacts and Vulnerabilities 27.3.6.2. Adaptation Practices 27.3.7. Human Health 27.3.7.1. Observed and Projected Impacts and Vulnerability 27.3.7.2. Adaptation Strategies and Practices 27.4. Adaptation Opportunities, Constraints, and Limits 27.4.1. Adaptation Needs and Gaps 27.4.2. Practical Experiences of Autonomous and Planned Adaptation, including Lessons Learned 27.4.3. Observed and Expected Barriers to Adaptation 27.5. Interactions between Adaptation and Mitigation 27.6. Case Studies 27.6.1. Hydropower 27.6.2. Payment for Ecosystem Services 27.7. Data and Research Gaps 27.8. Conclusions References Chapter Boxes 27-1. Extreme Events, Climate Change Perceptions, and Adaptive Capacity in Central America 27-2. Vulnerability of South American Megacities to Climate Change: The Case of the Metropolitan Region of Sao Paulo (MRSP) Frequently Asked Questions 27.1: What is the impact of glacier retreat on natural and human systems in the tropical Andes? 27.2: Can PES be used as an effective way for helping local communities to adapt to climate change? 27.3: Are there emerging and re emerging human diseases as a consequence of climate variability and change in the region? Executive Summary Significant trends in precipitation and temperature have been observed in Central America (CA) and South America (SA) (high confidence). Besides, changes in climate variability and in extreme events have severely affected the region (medium confidence). Increasing trends in annual rainfall in Southeastern South America (SESA; 0.6 mm/day/50years during 1950-2008) contrast with decreasing trends in CA and Central-Southern Chile (-1mm/day /50 years during 1950-2008). Warming has been detected throughout CA and SA (near to 0.7-1°C/40 years since the mid-1970 s), except for a cooling off the Chilean coast of about -1 C°/40 years. Increases in temperature extremes have been identified in CA and most of tropical and subtropical SA (medium confidence), while more frequent extreme rainfall in SESA has favoured the occurrence of landslides and flash floods (medium confidence). (27.2.1.1, Table 27-1, Box 27-1) Climate projections suggest increases in temperature, and increases or decreases in precipitation for CA and SA by 2100 (medium confidence). Post-AR4 climate projections, derived from dynamic downscaling forced by CMIP3 models for various SRES scenarios, and to different global climate models from the CMIP5 for various Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 RCPs (4.5 and 8.5), warming varies from +1.6°C to +4.0°C in CA, and +1.7°C to +6.7°C in SA (medium confidence). Rainfall changes for CA range between -22% to +7% by 2100, while in SA rainfall varies geographically, most notably showing a reduction of -22% in Northeast Brazil, and an increase of +25% in SESA (low confidence). By 2100 projections show an increase in dry spells in tropical SA east of the Andes, and in warm days and nights in most of SA (medium confidence) (27.2.1.2, Table 27-2). Changes in stream flow and water availability have been observed and projected to continue in the future in CA and SA, affecting already vulnerable regions (high confidence). The Andean cryosphere is retreating affecting the seasonal distribution of streamflows (Table 27-3) (high confidence). Increasing runoffs in the La Plata River basin and decreasing ones in the Central Andes (Chile, Argentina) and in CA in the second half of the 20th century were associated with changes in precipitation (high confidence). Risk of water supply shortages will increase owing to precipitation reductions and evapotranspiration increases in semi-arid regions (high confidence) (Table 27-4), thus affecting water supply for cities (high confidence) (27.3.1.1; 27.3.5), hydropower generation (high confidence) (27.3.6; 27.6.1) and agriculture (27.3.1.1). Current practices to reduce the mismatch between water supply and demand could be used to reduce future vulnerability (medium confidence). Ongoing constitutional and legal reforms towards more efficient and effective water resources management and coordination constitute another adaptation strategy (medium confidence) (27.3.1.2). Land use change contributes significantly to environmental degradation exacerbating the negative impacts of climate change (high confidence). Deforestation and land degradation are mainly attributed to increased extensive and intensive agriculture. The agricultural expansion, in some regions associated with increases in precipitation, has affected fragile ecosystems, such as the edges of the Amazon forest and the tropical Andes. Even though deforestation rates in the Amazon have decreased substantially since 2004 to a value of 4,656 km2/yr in 2012, other regions like the Cerrado still present high levels of deforestation with average rates as high as 14,179 km2/yr for the period 2002-2008 (27.2.2.1). Conversion of natural ecosystems is the main cause of biodiversity and ecosystem loss in the region, and is a driver of anthropogenic CC (high confidence). CC is expected to increase the rates of species extinction (medium confidence). For instance, vertebrate species turnover until 2100 will be as high as 90% in specific areas of CA and the Andes Mountains. In Brazil, distribution of some groups of birds and plants will be dislocated southwards, where there are less natural habitats remaining. However, CA and SA have still large extensions of natural vegetation cover for which the Amazon is the main example (27.3.2.1). Ecosystem-based adaptation practices are increasingly common across the region, such as the effective management and establishment of protected areas, conservation agreements and community management of natural areas (27.3.2.2). Socioeconomic conditions have improved since AR4; however there is still a high and persistent level of poverty in most countries resulting in high vulnerability and increasing risk to climate variability and change (high confidence). Poverty levels in most countries remain high (45% for CA and 30% for SA for year 2010) in spite of the sustained economic growth observed in the last decade. Human Development Index varies greatly between countries, from Chile and Argentina with the highest values, and Guatemala and Nicaragua with the lowest values in 2007. The economic inequality translates into inequality in access to water, sanitation and adequate housing, particularly for the most vulnerable groups translating into low adaptive capacities to climate change (27.2.2.2). Sea-level rise (SLR) and human activities on coastal and marine ecosystems pose threats to fish stocks, corals, mangroves, recreation and tourism, and control of diseases (high confidence). SLR varied from 2 to 7 mm/yr between 1950 and 2008. Frequent coral bleaching events associated to ocean warming and acidification occur in the Mesoamerican Coral Reef. In CA and SA, the main drivers of mangrove loss are deforestation and land conversion to agriculture and shrimp ponds (27.3.3.1). Brazilian fisheries co-management (a participatory multi-stakeholder process) is an example of adaptation since it favours a balance between conservation of marine biodiversity, the improvement of livelihoods, and the cultural survival of traditional populations (27.3.3.2). Changes in agricultural productivity with consequences for food security associated to CC are expected to exhibit large spatial variability (medium confidence). In SESA, where projections indicate more rainfall, average Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 productivity could be sustained or increased until the mid-century (SRES: A2, B2) (Table 27-5) (medium confidence). In CA, northeast of Brazil and parts of the Andean region increases in temperature and decreases in rainfall could decrease the productivity in the short-term (by 2030), threatening the food security of the poorest population (Table 27-5) (medium confidence). Considering that SA will be a key food producing region in the future, one of the challenges will be to increase the food and bioenergy quality and production while maintaining environmental sustainability under CC (27.3.4.1). Some adaptation measures include crop, risk, and water use management along with genetic improvement (27.3.4.2) (high confidence). Renewable energy (RE) based on biomass has a potential impact on land use change and deforestation and could be affected by CC (medium confidence). Sugarcane and soy are likely to respond positively to CO2 and temperature changes, even with a decrease in water availability, with an increase in productivity and production (high confidence). The expansion of sugarcane, soy and oil palm may have some effect on land use, leading to deforestation in parts of Amazon, CA among other regions, and loss of employment in some countries (medium confidence) (27.3.6.1). Advances in second-generation bioethanol from sugarcane and other feedstocks will be important as a measure of mitigation (27.3.6.2). Changes in weather and climatic patterns are negatively affecting human health in CA and SA, either by increasing morbidity, mortality, and disabilities (high confidence), and through the emergence of diseases in previously non-endemic areas (high confidence). With very high confidence climate-related drivers are associated with respiratory and cardiovascular diseases, vector- and water-borne diseases (malaria, dengue, yellow fever, leishmaniasis, cholera, and other diarrheal diseases), Hantaviruses and Rotaviruses, chronic kidney diseases, and psychological trauma. Air pollution is associated with pregnancy-related outcomes and diabetes, among others (27.3.7.1). Vulnerabilities vary with geography, age, gender, race, ethnicity, and socio-economic status, and are rising in large cities (27.3.7.2) (very high confidence). Climate change will exacerbate current and future risks to health, given the region s population growth rates and vulnerabilities in existing health, water, sanitation and waste collection systems, nutrition, pollution and food production in poor regions (medium confidence). In many CA and SA countries, a first step toward adaptation to future climate changes is to reduce the vulnerability to present climate. Long-term planning and the related human and financial resource needs may be seen as conflicting with present social deficit in the welfare of the CA and SA population. Various examples demonstrate possible synergies between development, adaptation and mitigation planning, which can help local communities and governments to allocate efficiently available resources in the design of strategies to reduce vulnerability. However, the generalization of such actions at continental scale requires that both the CA and SA citizens and governments are faced with the challenge of building a new governance model, where imperative development needs, vulnerability reduction and adaptation strategies to climate stresses will be truly intertwined. (27.3.4, 27.4.1, 27.4.2, 27.4.3, 27.4.4, 27.5). 27.1. Introduction 27.1.1. The Central and South America Region The CA and SA region harbors unique ecosystems, the highest biodiversity in the planet and has a variety of eco- climatic gradients. Unfortunately, this natural wealth is threatened by advancing agricultural frontiers resulting from a rapidly growing agricultural and cattle production (Grau and Aide, 2008). The region experienced a steady economic growth in the last decade, accelerated urbanization and important demographic changes; poverty and inequality are decreasing continuously, but at a low pace (ECLAC, 2011c). Adaptive capacity is improving in part thanks to poverty alleviation and development initiatives (McGray et al., 2007). The region has multiple stressors on natural and human systems derived in part from significant land use changes and exacerbated by climate variability/climate change. Climate variability at various time scales has been affecting social and natural systems, and extremes in particular have affected large regions. During 2000-2010, almost 630 weather and climate extreme events occurred in CA and SA, leaving near to 16,000 fatalities and 46.6 million Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 people affected; and generating economical losses amounting to US$ 208 million (CRED, 2011). Land is facing increasing pressure from competing uses like cattle ranching, food production and bioenergy. The region is regarded as playing a key role in future world economy because countries like Brazil, Chile, Colombia and Panama, among others, are rapidly developing and becoming economically important in the world scenario. The region is bound to be exposed to the pressure related to increasing land use and industrialization. Therefore, it is expected to have to deal with increasing emission potentials. Thus, science-based decision-making is thought to be an important tool to control innovation and development of the countries in the region. Two other important contrasting features characterize the region: having the biggest tropical forest of the planet on the one side, and possessing the largest potential for agricultural expansion and development during the next decades on the other. This is the case because the large countries of SA, especially, would have a major role in food and bioenergy production in the future, as long as policies towards adaptation to global climate change (GCC) will be strategically designed. The region is already one of the top producers and user of bioenergy and this experience will serve as an example to other developing regions as well as developed regions. 27.1.2. Summary of the AR4 and SREX Findings 27.1.2.1. AR4 Findings According to AR4-Chapter 13 (Latin America), during the last decades of the 20th century, unusual extreme weather events have been severely affecting the LA region contributing greatly to the strengthening of the vulnerability of human systems to natural disasters. In addition, increases in precipitation were observed in SESA, northwest Peru and Ecuador; while decreases were registered in southern Chile, southwest Argentina, southern Peru and western CA since 1960. Mean warming was near to 0.1C/decade. The rate of SLR has accelerated over the last 20 years reaching 2-3mm/year. The glacier-retreat trend has intensified, reaching critical conditions in the Andean countries. Rates of deforestation have been continuously increasing mainly due to agricultural expansion, and land degradation has been intensified for the entire region. Mean warming for LA at the end of 21st century could reach 1C to 4C (SRES B2) or 2C to 6C (SRES A2) (medium confidence) (AR4-Ch13-pp583). Rainfall anomalies (positive or negative) will be larger for the tropical part of LA. The frequency and intensity of weather and climate extremes is likely to increase (medium confidence). Future impacts include: Significant species extinctions, mainly in tropical LA (high confidence). Replacement of tropical forest by savannas, and semi-arid vegetation by arid vegetation (medium confidence). Increases in the number of people experiencing water stress (medium confidence). Probable reductions in rice yields and possible increases of soy yield in SESA (AR4-Ch13-pp583); and increases in crop pests and diseases (medium confidence) (AR4-Ch13-pp607). Some coastal areas affected by sea level rise, weather and climatic variability and extremes (high confidence) (AR4-Ch13-pp584). Some countries have made efforts to adapt to climate change and variability, for example through the conservation of key ecosystems (e.g. biological corridors in Mesoamerica, Amazonia, and Atlantic forest; compensation for ecosystem services in Costa Rica,), the use of early warning systems and climate forecast (e.g. fisheries in eastern Pacific, subsistence agriculture in NE Brazil), and the implementation of disease surveillance systems (e.g. Colombia) (AR4-Ch13-pp 591). However, several constraints like the lack of basic information, observation and monitoring systems; the lack of capacity-building and appropriate political, institutional and technological frameworks; low income; and settlements in vulnerable areas, outweigh the effectiveness of these efforts. 27.1.2.2. SREX Findings As reported on Chapter 3.4 of the IPCC Special Report on Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation (SREX) (IPCC, 2012b), a changing climate leads to changes in the frequency, Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 intensity, spatial extent or duration of weather and climate extremes, and can result in unprecedented extremes. Levels of confidence in historical changes depend on the availability of high quality and homogeneous data, and relevant model projections. This has been a major problem in CA and SA, where a lack of long-term homogeneous and continuous climate and hydrological records, and of complete studies on trends have not allowed for an identification of trends in extremes, particularly in CA. Recent observational studies and projections from global and regional models suggest changes in extremes. With medium confidence, increases in warm days and decreases in cold days, as well as increases on warm nights and decreases in cold nights have been identified in CA, Northern SA, NEB, SESA and west coast of SA. In CA, there is low confidence that any observed long-term increase in tropical cyclone activity is robust, after accounting for past changes in observing capabilities. In other regions, such as the Amazon region, insufficient evidence, inconsistencies among studies and detected trends result in low confidence of observed rainfall trends. While it is likely that there has been an anthropogenic influence on extreme temperature in the region, there is low confidence in attribution of changes in tropical cyclone activity to anthropogenic influences. Projections for the end of the 21st century for differing emissions scenarios (SRES A2 and A1B) show that for all CA and SA, models project substantial warming in temperature extremes. It is likely that increases in the frequency and magnitude of warm daily temperature extremes and decreases in cold extremes will occur in the 21st century on the global scale. With medium confidence, it is very likely that the length, frequency and/or intensity of heat waves will experience a large increase over most of SA, with weaker tendency towards increasing in SESA. With low confidence, the models also project an increase of the proportion of total rainfall from heavy falls for SESA and the West coast of SA; while for Amazonia and the rest of SA and CA there are not consistent signals of change. In some regions, there is low confidence in projections of changes in fluvial floods. Confidence is low due to limited evidence and because the causes of regional changes are complex. There is medium confidence that droughts will intensify along the 21st century in some seasons and areas due to reduced precipitation and/or increased evapotranspiration in Amazonia and NEB. The character and severity of the impacts from climate extremes depend not only on the extremes themselves but also on exposure and vulnerability. These are influenced by a wide range of factors, including anthropogenic climate change, natural climate variability, and socioeconomic development. 27.2. Major Recent Changes and Projections in the Region 27.2.1. Climatic Stressors 27.2.1.1. Climate Trends, Long-term Changes in Variability, and Extremes In CA and SA, decadal variability and changes in extremes have been affecting large sectors of the population, especially those more vulnerable and exposed to climate hazards. Observed changes in some regions have been attributed to natural climate variability, while in other regions they have been attributed to land use change, e.g. increase urbanization, meaning that land use change is a result of anthropogenic drivers. Table 27-1 summarizes observed trends in the region s climate. [INSERT TABLE 27-1 HERE Table 27-1: Regional observed changes in temperature, precipitation and climate extremes in various sectors of CA and SA. Additional information on changes in observed extremes can be found in the IPCC SREX (Seneviratne et al., 2012) and Chapter 2 IPCC WGI AR5 [2.4, 2.5, 2.6]] Since around 1950, in CA and the North American Monsoon System (NAMS), rainfall has been starting increasingly later and has become more irregular in space and time, while rainfall has been increasing and the intensity of rainfall has been increasing during the onset season (see references in Table 27-1). Arias et al. (2012) relate those changes to decadal rainfall variations in NAMS. Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The West coast of SA experienced a prominent but localized coastal cooling of about 1 C during the past 30-50 years extending from central Peru down to central Chile. This occurs in connection with an increased upwelling of coastal waters favored by the more intense trade winds (Falvey and Garreaud, 2009; Gutiérrez et al., 2011a; Gutiérrez et al., 2011b; Kosaka and Xie, 2013; Narayan et al., 2010; Schulz et al., 2012). In the extremely arid northern coast of Chile, rainfall, temperature and cloudiness show strong interannual and decadal variability, and since the mid-70s, the minimum daily temperature, cloudiness and precipitation have decreased. In central Chile, a negative precipitation trend was observed over the period 1935-1976, and an increase after 1976, while further south, the negative trend in rainfall that prevailed since the 1950s has intensified by the end of the 20th century (Quintana and Aceituno, 2012). To the east of the Andes, northeast Brazil (NEB) exhibits large interannual rainfall variability, with a slight decrease since the 1970s (Marengo et al. 2013a). Droughts in this region (e.g. 1983, 1987, 1998) have been associated with El Nino and/or a warmer Tropical North Atlantic Ocean. However, not all El Nino years result on drought in NEB, as the drought 2012-2013 occurred during La Nina (Marengo et al., 2013a). In the La Plata Basin in SESA, various studies have documented interannual and decadal scale circulation changes that have led to decreases in the frequency of cold nights in austral summer, as well as to increases in warm nights and minimum temperatures during the last 40 years. Simultaneously, a reduction in the number of dry months in the warm season is found since the mid-1970s, while heavy rain frequency is increasing in SESA (references in Table 27-1). In SESA, increases in precipitation are responsible for changes in soil moisture (Collini et al., 2008; Saulo et al., 2010), and although feedback mechanisms are present at all scales, the effect on atmospheric circulation is detected at large scales. Moreover, land use change studies in the Brazilian southern Amazonia for the last decades showed that the impact on the hydrological response is time lagged at larger scales (Rodriguez et al., 2010) In the central Andes, in the Mantaro Valley (Peru), precipitation shows a strong negative trend, while warming is also detected (SENAMHI, 2007). In the southern Andes of Peru air temperatures have increased during 1964-2006, but no clear signal on precipitation changes has been detected (Marengo et al., 2009a). In the northern Andes (Colombia, Ecuador), changes in temperature and rainfall in 1961-90 have been identified by Villacís (2008). In the Patagonia region, Masiokas et al. (2008) have identified an increase of temperature together with precipitation reductions during 1912-2002. Vuille et al. (2008a) found that climate in the tropical Andes has changed significantly over the past 50 60 years. Temperature in the Andes has increased by approximately 0.1 °C/ decade, with only two of the last 20 years being below the 1961 90 average. Precipitation has slightly increased in the second half of the 20th century in the inner tropics and decreased in the outer tropics. The general pattern of moistening in the inner tropics and drying in the subtropical Andes is dynamically consistent with observed changes in the large-scale circulation, suggesting a strengthening of the tropical atmospheric circulation. Moreover, a positive significant trend in mean temperature of 0.09 oC per decade during 1965-2007 has been detected over the Peruvian Andes by Lavado et al. (2012). For the Amazon basin, Marengo (2004) and Satyamurty et al. (2010) concluded that no systematic unidirectional long-term trends towards drier or wetter conditions in both the northern and southern Amazon have been identified since the 1920s. Rainfall fluctuations are more characterized by inter-annual scales linked to ENSO or decadal variability. Analyzing a narrower time period, Espinoza et al. (2009a; 2009b) found that mean rainfall in the Amazon basin for 1964 2003 has decreased, with stronger amplitude after 1982, especially in the Peruvian western Amazonia (Lavado et al., 2012), consistent with reductions in convection and cloudiness in the same region (Arias et al., 2011). Recent studies by Donat et al. (2013) suggest that heavy rains are increasing in frequency in Amazonia. Regarding seasonal extremes in the Amazon region, two major droughts and three floods have affected the region from 2005 to 2012, although these events have been related to natural climate variability rather than to deforestation (Espinoza et al., 2013; Espinoza et al., 2011; Espinoza et al., 2012; Lewis et al., 2011; Marengo et al., 2008; Marengo et al., 2012; Marengo et al., 2013a; Satyamurty et al., 2013). On the impacts of land use changes on changes in the climate and hydrology of Amazonia, Zhang et al. (2009) suggest that biomass-burning aerosols can work against the seasonal monsoon circulation transition, thus re-inforce the dry season rainfall pattern for Southern Amazonia, while Wang et al. (2011) suggests the importance of deforestation and vegetation dynamics on decadal variability of rainfall in the region. Costa and Pires (2010) have suggested a possible decrease in precipitation due to soybean expansion in Amazonia, mainly as a consequence of its very high albedo. In the SAMS region, positive Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 trends in rainfall extremes have been identified in the last 30 years with a pattern of increasing frequency and intensity of heavy rainfall events, and earlier onsets and late demise of the rainy season (see Table 27-1). ______START BOX 27-1 HERE______ Box 27-1. Extreme Events, Climate Change Perceptions, and Adaptive Capacity in Central America Central America has traditionally been characterized as a region with high exposure to geo-climatic hazards derived from its location and topography and with high vulnerability of its human settlements (ECLAC, 2010c). It has also been identified as the most responsive tropical region to climate change (Giorgi, 2006). Evidence for this has been accumulating particularly in the last 30 years with a steady increase in extreme events including storms, floods and droughts. In the period 2000-2009, 39 hurricanes occurred in the Caribbean basin compared to 15 and 9 in the decade of 1980 and 1990 respectively (UNEP-ECLAC, 2010). The impacts of these events on the population and the economy of the region have been tremendous: the economic loss derived from 11 recent hydrometereological events evaluated added to US $13,642 millions and the number of people impacted peaked with hurricane Mitch in 1998 with over 600,000 persons affected (ECLAC, 2010c). A high percentage of the population in CA live on or near highly unstable steep terrain with sandy, volcanic soils prone to mudslides which are the main cause of casualties and destruction (Restrepo and Alvarez, 2006). The increased climatic variability in the past decade certainly changed the perception of people in the region with respect to climate change. In a survey to small farmers in 2003, Tucker et al. (2010) found that only 25% of respondents included climate events as a major concern. A subsequent survey in 2007 (Eakin et al., 2013) found that over 50% of respondents cited drought conditions and torrential rains as their greatest concern. Interestingly, there was no consensus on the direction in climate change pattern: the majority or households in Honduras reported an increase in the frequency of droughts but in Costa Rica and Guatemala a decrease or no trend at all was reported; a similar discrepancy in answers was reported with the issue of increase rainfall. But there was general agreement in all countries that rainfall patterns were more variable resulting in higher difficulty in recognizing the start of the rainy season. The high levels of risk to disasters in CA are the result of high exposure to hazards and the high vulnerability of the population and its livelihoods derived from elevated levels of poverty and social exclusion (Programa Estado de la Nación-Región, 2011). Disaster management in the region has focused on improving early warning systems and emergency response for specific extreme events (Saldana-Zorrilla, 2008) but little attention has been paid to strengthening existing social capital in the form of local organizations and cooperatives. These associations can be central in increasing adaptive capacity through increased access to financial instruments and strategic information on global markets and climate (Eakin et al., 2011). There is a need to increase the communication of the knowledge from local communities involved in processes of autonomous adaptation to policy makers responsible for strengthening the adaptive capacities in CA (Castellanos et al., 2013). ______END BOX 27-1 HERE______ 27.2.1.2. Climate Projections Since the AR4, substantial additional regional analysis has been carried out using the CMIP3 model ensemble. In addition, projections from CMIP5 models and new experiences using regional models (downscaling) have allowed for a better description of future changes in climate and extremes in CA and SA. Using CMIP3 and CMIP5 models, Giorgi (2006), Diffenbaugh et al. (2008), Xu et al. (2009), Diffenbaugh and Giorgi (2012) and Jones and Carvalho (2013) have identified areas of CA/western North America and the Amazon as persistent regional climate change hotspots throughout the 21st century of the RCP8.5 and RCP4.5. Table 27-2 summarizes projected climatic changes derived from global and regional models for the region, indicating the projected change, models, emission scenarios, time spans and references. Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT TABLE 27-2 HERE Table 27-2: Regional projected changes in temperature, precipitation, and climate extremes in different sectors of CA and SA. Various studies used A2 and B2 scenarios from CMIP3 and various RCPs scenarios for CMIP5, and different time slices from 2010 to 2100. In order to make results comparable, the CMIP3 and CMIP5 at the time slice ending in 2100 are included. Additional information on changes in projected extremes can be found in the IPCC SREX (see IPCC, 2012b), and Chapters 9 and 14 from IPCC WG1AR5 [9.5, 9.6 and 14.2, 14.7]] In CA and Northern Venezuela, projections from CMIP3 models and from downscaling experiments suggest precipitation reductions and warming together with an increase in evaporation, and reductions in soil moisture for most of the land during all seasons by the end of the 21th century (see references in Table 27-2). However, the spread of projections is high for future precipitation. Analyses from global and regional models in tropical and subtropical SA show common patterns of projected climate in some sectors of the continent. Projections from CMIP3 regional and high resolution global models show by the end of the 21st century for the A2 emission scenario, a consistent pattern of increase of precipitation in SESA, Northwest of Peru and Ecuador and western Amazonia, while decreases are projected for northern SA, Eastern Amazonia, central eastern Brazil, NEB, the Altiplano and southern Chile (Table 27-2). For some regions, projections show mixed results in rainfall projections for the Amazonia and the SAMS region suggesting high uncertainties on the projections (Table 27-2). As for extremes, CMIP3 models and downscaling experiments show increases in dry spells are projected for Eastern Amazonia and NEB, while rainfall extremes are projected to increase in SESA, in western Amazonia, Northwest Peru and Ecuador, while over southern Amazonia, northeastern Brazil and eastern Amazonia, the maximum number of consecutive dry days tends to augment, suggesting a longer dry season. Increases in warm nights throughout SA are also projected by the end of the 21st century (see references in Table 27-2). Shiogama et al. (2011) suggest that although the CMIP3 ensemble mean assessment suggested wetting across most of SA, the observational constraints indicate a higher probability of drying in the eastern Amazon basin. The CMIP5 models project an even larger expansion of the monsoon regions in NAMS in future scenarios (Jones and Carvalho, 2013; Kitoh et al., 2013). A comparison from eight models from CMIP3 and CMIP5 identifies some improvements in the new generation models. For example, CMIP5 inter-model variability of temperature in summer was lower over northeastern Argentina, Paraguay and northern Brazil, in the last decades of the 21st century, as compared to CMIP3. Although no major differences were observed in both precipitation datasets, CMIP5 inter- model variability was lower over northern and eastern Brazil in summer by 2100 (Blázquez and Nunez, 2013; Jones and Carvalho, 2013). The projections from the CMIP5 models at regional level for CA and SA (using the same regions from the IPCC SREX) are shown in Figure 27-1, and update some of these previous projections based on SRES A2 and B2 emission scenarios from CMIP3. Figure 27-1 shows that in relation to the baseline period 1986-2005, for CA and northern South America-Amazonia, temperatures are projected to increase approximately by 0.6 °C and 2 °C for the RCP2.6 scenario, and by 3.6 C and 5.2 °C for the RCP8.5 scenario. For the rest of South America, increases by about 0.6 °C to 2 °C are projected for the RCP4.5 and by about 2.2 °C to 7 °C for the RCP8.5 scenario. The observed records show increases of temperature from 1900 to 1986 by about 1 °C. For precipitation, while for CA and northern South America-Amazonia precipitation is projected to vary between +10% to -25% (with large spread among models). For NEB, there is a spread among models between +30 to -30% making hard to identify any projected rainfall change. This spread is much lower in the western coast of South America and SESA, where the spread is between +20 and -10% (IPCC WG2 Chapter 21, Box 21-3). Figure 27-2 shows that by late-21st century, the CMIP5 derived projections for RCP8.5 projected: CA: mean annual warming of 2.5C and rainfall reduction of 10%, and reduction in summertime precipitation. SA: mean warming of 4C, with rainfall reduction up to 15% in tropical SA east of the Andes, and an increase of about 15-20% in SESA, and in other regions of the continent. Changes in mid-21st century are small. Both Figures 27-1 and 27-2 show that there is some degree of uncertainty on climate change projections for regions, particularly for rainfall in CA and tropical SA. Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE 27-1 HERE Figure 27-1: Observed and simulated variations in past and projected future annual average temperature over the Central and South American regions defined in IPCC (2012a). Black lines show various estimates from observational measurements. Shading denotes the 5-95 percentile range of climate model simulations driven with "historical" changes in anthropogenic and natural drivers (63 simulations), historical changes in "natural" drivers only (34), the "RCP2.6" emissions scenario (63), and the "RCP8.5" (63). Data are anomalies from the 1986-2006 average of the individual observational data (for the observational time series) or of the corresponding historical all- forcing simulations. Further details are given in Box 21-3.] [INSERT FIGURE 27-2 HERE Figure 27-2: Projected changes in annual average temperature and precipitation. CMIP5 multi-model mean projections of annual average temperature changes (left panel) and average percent change in annual mean precipitation (right panel) for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and>90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where>66% of models show change greater than the baseline variability and>66% of models agree on sign of change. Gray indicates areas with divergent changes, where>66% of models show change greater than the baseline variability, but<66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in>66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5. [Boxes 21-3 and CC-RC]] 27.2.2. Non-Climatic Stressors 27.2.2.1. Trends and Projections in Land Use and Land Use Change Land use change is a key driver of environmental degradation for the region that exacerbates the negative impacts from climate change (Lopez-Rodriguez and Blanco-Libreros, 2008; Sampaio et al., 2007). The high levels of deforestation observed in most of the countries in the region have been widely discussed in the literature as a deliberate development strategy based on the expansion of agriculture to satisfy the growing world demand for food, energy and minerals (Benhin, 2006; Grau and Aide, 2008; Mueller et al., 2008). Land is facing increasing pressure from competing uses, among them cattle ranching, food and bioenergy production. The enhanced competition for land increases the risk of land use changes, which may lead to negative environmental and socio-economic impacts. Agricultural expansion has relied in many cases on government subsidies, which have often resulted in lower land productivity and more land speculation (Bulte et al., 2007; Roebeling and Hendrix, 2010). Some of the most affected areas due to the expansion of the agricultural frontier are fragile ecosystems such as the edges of the Amazon forest in Brazil, Colombia, Ecuador and Peru, and the tropical Andes including the Paramo, where activities such as deforestation, agriculture, cattle ranching and gold mining are causing severe environmental degradation (ECLAC, 2010d), and the reduction of environmental services provided by these ecosystems. Deforestation rates for the region remain high in spite of a reducing trend in the last decade (Fearnside, 2008; Ramankutty et al., 2007). Brazil is by far the country with the highest area of forest loss in the world according to the latest FAO statistics (2010): 21,940 km per year equivalent to 39% of the world deforestation for the period 2005-2010. Bolivia, Venezuela and Argentina follow in deforested area (Figure 27-3) with 5.5%, 5.2% and 4.3% of the total world deforestation respectively. The countries of CA and SA lost a total of 38,300 km of forest per year in that period (69% of the total world deforestation) (FAO, 2010). These numbers are limited by the fact that many countries do not have comparable information through time, particularly for recent years. Aide et al. (2013) completed a wall-to-wall analysis for the region for the period 2001-2010 analyzing not only deforestation but also reforestation and reported very different results than FAO (2010) for some countries where reforestation seems to be higher than deforestation, particularly in Honduras, El Salvador, Panama, Colombia and Venezuela. For Colombia and Venezuela, these results are contradictory with country analyses that align better with the FAO data (Armenteras et al., 2013; Rodríguez et al., 2010). Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE 27-3 HERE Figure 27-3: Area deforested per year for selected countries in CA and SA (2005-2010). Notice three countries listed with a positive change in forest cover (based on data from FAO, 2010).] Deforestation in the Amazon forest has received much international attention in the last decades, both because of its high rates, and its rich biodiversity. Brazilian Legal Amazon is now one of the best-monitored ecosystems in terms of deforestation since 1988 (INPE, 2011). Deforestation for this region peaked in 2004 and has steadily declined since then to a lowest value of 4,656 Km2/yr for the year 2012 (see Figure 27-4). Such reduction results from a series of integrated policies to control illegal deforestation particularly enforcing protected areas, which now shelter 54% of the remaining forests of the Brazilian Amazon (Soares-Filho et al., 2010). Deforestation in Brazil is now highest in the Cerrado (drier ecosystem south of Amazon) with an average value of 14,179 Km2/yr for the period 2002-2008 (FAO, 2009). [INSERT FIGURE 27-4 HERE Figure 27-4: Deforestation rates in the Brazilian Amazonia (km/year) based on measurements by the PRODES project (INPE, 2011).] The area of forest loss in CA is considerably less than in SA, owing to smaller country sizes (Carr et al., 2009), but when relative deforestation rates are considered, Honduras and Nicaragua show the highest values for CA and SA (FAO, 2010). At the same time, CA includes some countries where forest cover shows a small recovery trend in the last years: Costa Rica, El Salvador, Panama and possibly Honduras where data is conflicting in the literature (Aide et al., 2013; FAO, 2010). This forest transition is the result of: (1) economies less dependent on agriculture, and more on industry and services (Wright and Samaniego, 2008); (2) processes of international migration with the associated remittances (Hecht and Saatchi, 2007), and (3) a stronger emphasis on the recognition of environmental services of forest ecosystems (Kaimowitz, 2008). The same positive trend is observed in some SA countries (Figure 27-3). However, a substantial amount of forest is gained through (single-crop) plantations, most noticeably in Chile (Aguayo et al., 2009), which has a much lower ecological value than the depleted natural forests (Echeverría et al., 2006; Izquierdo et al., 2008). Land degradation, is also an important process compromising extensive areas of CA and SA very rapidly. According to data from the Global Land Degradation Assessment and Improvement (GLADA) project of the Global Environmental Facility (GEF), additional degraded areas reached 16.4% of the entire territory of Paraguay, 15.3% of Peru and 14.2% of Ecuador for the period 1982-2002. In CA, Guatemala shows the highest proportion of degraded land, currently at 58.9% of the country s territory, followed by Honduras (38.4%) and Costa Rica (29.5%); only El Salvador shows a reversal of the land degradation process, probably due to eased land exploitation following intensive international migratory processes (ECLAC, 2010d). Deforestation and land degradation are mainly attributed to increased extensive and intensive agriculture. Two activities have traditionally dominated the agricultural expansion: soy production (only in SA) and beef; but more recently, biomass for biofuel production has become as important (Nepstad and Stickler, 2008) with some regions also affected by oil and mining extractions. Deforestation by small farmers, mainly coming from families who migrate in search for land is relatively low: extensive cattle production is the predominant land use in deforested areas of tropical and subtropical Latin America (Wassenaar et al., 2007). Cattle is the only land use variable correlated with deforestation in Colombia (Armenteras et al., 2013) and in the Brazilian Amazon the peak of deforestation in 2004 (Figure 27-4) was primarily the result of increased cattle ranching (Nepstad et al., 2006). Mechanized farming, agro-industrial production and cattle ranching are the major land use change drivers in eastern Bolivia but subsistence agriculture by indigenous colonists is also important (Killeen et al., 2008). In recent years, soybean croplands have expanded continuously in SA, becoming increasingly more important in the agricultural production of the region. Soybean-planted area in Amazonian states (mainly Mato Grosso) in Brazil expanded 12.1% per year during the 1990s, and 16.8% per year from 2000 to 2005 (Costa et al., 2007). This landscape-scale conversion from forest to soy and other large-scale agriculture can alter substantially the water Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 balance for large areas of the region resulting in important feedbacks to the local climate (Hayhoe et al., 2011; Loarie et al., 2011) (see section 27.3.4.1). Soybean and beef production have also impacted other ecosystems next to the Amazon, such as the Cerrado (Brazil) and the Chaco dry forests (Bolivia, Paraguay, Argentina and Brazil). Gasparri et al. (2008) estimated carbon emissions from deforestation in Northern Argentina, and concluded that deforestation in the Chaco forest has accelerated in the past decade from agricultural expansion and is now the most important source of carbon emission for that region. In northwest Argentina (Tucumán and Salta provinces) 14,000 km 2 of dry forest were cleared from 1972 to 2007 as a result of technological improvements and increasing rainfall (Gasparri and Grau, 2009). Deforestation continued during the 1980s and 1990s resulting in cropland area covering up to 63% of the region by 2005 (Viglizzo et al., 2011). In central Argentina (northern Córdoba province), cultivated lands have increased from 3% to 30% (between 1969 and 1999); and the forest cover has decreased from 52.5% to 8.2%. This change has also been attributed to the synergistic effect of climatic, socioeconomic, and technological factors (Zak et al., 2008). Losses in the Atlantic forest are estimated in 29% of the original area in 1960, and in 28% of the Yunga forest area, mainly due to cattle ranching migration from the Pampas and Espinal (Viglizzo et al., 2011). Oil palm is a significant biofuel crop also linked to recent deforestation in tropical CA and SA. Its magnitude is still small compared with deforestation related to soybean and cattle ranching, but it is considerable for specific countries and expected to increase due to increasing demands for biofuels (Fitzherbert et al., 2008). The main producers of palm oil in the region are Colombia and Ecuador, followed by Costa Rica, Honduras, Guatemala and Brazil; this country has the largest potential for expansion, as nearly half of the Amazonia is suitable for oil palm cultivation (Butler and Laurance, 2009). Oil palm production is also growing in the Amazonian region of Peru, where 72% of new plantations have expanded into forested areas representing 1.3% of the total deforestation for that country for the years 2000-2010 (Gutiérrez-Vélez et al., 2011). However, forests are not the only important ecosystems threatened in the region. An assessment of threatened ecosystems in SA by Jarvis et al. (2010) concluded that grasslands, savannas and shrublands are more threatened than forests, mainly from excessively frequent fires (>1/year) and grazing pressure. An estimation of burned land in Latin America by Chuvieco et al. (2008) also concluded that herbaceous areas presented the highest occurrence of fires. In the Río de la Plata region (Central-East Argentina, southern Brazil, and Uruguay), grasslands decreased from 67.4% to 61.4% between 1985 and 2004. This reduction was associated with an increase in annual crops, mainly soybean, sunflower, wheat, and maize (Baldi and Paruelo, 2008). Even with technological changes that might result in agricultural intensification, the expansion of pastures and croplands is expected to continue in the coming years (Kaimowitz and Angelsen, 2008; Wassenaar et al., 2007), particularly from an increasing global demand for food and biofuels (Gregg and Smith, 2010) with the consequent increase in commodity prices. This agricultural expansion will be mainly in Latin America and Sub-Saharan Africa as these regions hold two-thirds of the global land with potential to expand cultivation (Nepstad and Stickler, 2008). It is important to consider the policy and legal needs to keep this process of large-scale change under control as much as possible; Takasaki (2007) showed that policies to eliminate land price distortions and promote technological transfers to poor colonists could reduce deforestation. It is also important to consider the role of indigenous groups; there is a growing acknowledgment that recognizing the land ownership and authority of indigenous groups can help central governments to better manage many of the natural areas remaining in the region (Larson, 2010; Oltremari and Jackson, 2006). The impact of indigenous groups on land use change can vary: Oliveira et al. (2007) found that only 9% of the deforestation in the Peruvian Amazon between 1999 and 2005 happened in indigenous territories but Killeen et al. (2008) found that Andean indigenous colonists in Bolivia were responsible for the largest land cover changes in the period 2001-2004. Indigenous groups are important stakeholders in many territories in the region and their well-being should be considered when designing responses to pressures on the land by a globalized economy (Gray et al., 2008; Killeen et al., 2008). Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 27.2.2.2. Trends and Projections in Socioeconomic Conditions Development in the region has traditionally displayed four characteristics: low growth rates, high volatility, structural heterogeneity and a very unequal income distribution (Bárcena, 2010; ECLAC, 2008). This combination of factors has generated high and persistent poverty levels (45% for CA and 30% for SA for year 2010), with the rate of poverty being generally higher in rural than urban areas (ECLAC, 2009b). SA has based its economic growth in natural resource exploitation (mining, energy, agricultural), which involves direct and intensive use of land and water, and in energy-intensive and, in many cases, highly polluting natural-resource-based manufactures. In turn, CA has exploited its proximity to the North American market and its relatively low labor costs (ECLAC, 2010e). The region shows a marked structural heterogeneity, where modern production structures coexist with large segments of the population with low productivity and income levels (ECLAC, 2010g). The GDP per capita in SA is twice that of CA; in addition, in the latter poverty is 50% higher (see Figure 27-5). [INSERT FIGURE 27-5 HERE Figure 27-5: Evolution of GDP per capita and poverty (income below US$ 2 per day) from 1990-2010: CA and SA (US-Dollars per inhabitant at 2005 prices and percentages) (ECLAC on the basis of CEPALSTAT (2012) and ECLAC (2011c)).] The 2008 financial crisis reached CA and SA through exports and credits, remittances and worsening expectations by consumers and producers (Bárcena, 2010; Kacef and López-Monti, 2010). This resulted in the sudden stop of six consecutive years of robust growth and improving social indicators (ECLAC, 2010e), which contributed to higher poverty in 2009 after six years where poverty had declined by 11%. Poverty rates fell from 44% to 33% of the total population from 2003 to 2008 (Figure 27-5), leaving 150 million people in this situation while extreme poverty diminished from 19.4% to 12.9% (which represents slightly more than 70 million people) (ECLAC, 2009b). In the second half of 2009 industrial production and exports began to recover and yielded a stronger economic performance (GDP growth of 6.4% in SA and 3.9% in CA in 2010) (ECLAC, 2012). SA benefited the most because of the larger size of their domestic markets and the greater diversification of export markets. Conversely, slower growth was observed in CA with more open economies and a less diversified portfolio of trading partners and a greater emphasis on manufacturing trade (ECLAC, 2010g). The region is expected to continue to grow in the short term, albeit at a pace that is closer to potential GDP growth, helped by internal demand as the middle class becomes stronger and as credit becomes more available. In SA, this could be boosted by external demand from the Asian economies as they continue to grow at a rapid pace. The macroeconomic challenge is to act counter cyclically creating conditions for productive development that is not based solely on commodity exports (ECLAC, 2010f). In spite of its economic growth, CA and SA still display high and persistent inequality: most countries have Gini coefficients between 0.5 and 0.6, whereas the equivalent figures in a group of 24 developed countries vary between under 0.25 and around 0.4. The average per capita income of richest 10% of households is approximately 17 times that of the poorest 40% of households (ECLAC, 2010g). Nevertheless, during the first decade of the century, prior to the financial crisis, the region has shown a slight but clear trend towards a more equitable distribution of income and a stronger middle class population resulting in a higher demand for goods (ECLAC, 2010g; ECLAC, 2011b; UN, 2010). Latin American countries also reported gains in terms of human development, although these gains have slowed down slightly over recent years. In comparative terms, the performance of countries as measured by the Human Development Index (HDI) varied greatly in 2007 (from Chile with 0.878 and Argentina with 0.866 to Guatemala with 0.704 and Nicaragua with 0.699) although those with lower levels of HDI showed notably higher improvements than countries with the highest HDI (UNDP, 2010). Associated with inequality are disparities in access to water, sanitation and adequate housing for the most vulnerable groups - for example indigenous peoples, Afro-descendants, children and women living in poverty- and in their exposure to the effects of climate change. The strong heterogeneity of subnational territorial entities in the region takes the form of high spatial concentration and persistent disparities in the territorial distribution of wealth (ECLAC, 2010g; ECLAC, 2011b; UN, 2010). Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The region faces significant challenges in terms of environmental sustainability and adaptability to a changing climate (UN, 2010), resulting from the specific characteristics of its population and economy already discussed and aggravated with a significant deficit in infrastructure development. The countries in CA and SA have made progress in incorporating environmental protection into decision-making processes, and particularly in terms of environmental institutions and legislation, but there are still difficulties to effectively incorporate environmental issues into relevant public policies (UN, 2010). While climate change imposes new challenges, it also provides an opportunity to shift development and economic growth patterns towards a more environmentally friendly course. 27.3. Impacts, Vulnerabilities, and Adaptation Practices 27.3.1. Freshwater Resources CA and SA are regions with a high average but unevenly distributed water resources availability (Magrin et al., 2007a). The main user of water is agriculture followed by the region s 580 million inhabitants (including the Caribbean), of which 86% had access to water supply by 2006 (ECLAC, 2010b). According to the International Energy Agency (IEA), the region meets 60% of its electricity demand through hydropower generation, which contrast with the 20% average contribution of other regions (see Table 27-6 and case study 27.6.1). 27.3.1.1. Observed and Projected Impacts and Vulnerabilities In CA and SA there are many evidences of changing hydrologic related conditions. The most robust trend for major rivers is found in the sub-basins of the La Plata River basin (high confidence based on high agreement, robust evidence). This basin, second only to the Amazon in size, shows a positive trend in streamflow in the second half of the 20th century at different sites (Amsler and Drago, 2009; Conway and Mahé, 2009; Dai et al., 2009; Dai, 2011; Doyle and Barros, 2011; Krepper et al., 2008; Krepper and Zucarelli, 2010a; Pasquini and Depetris, 2007; Saurral et al., 2008). An increase in precipitation and a reduction in evapotranspiration from land use changes have been associated with the trend in streamflows (Doyle and Barros, 2011; Saurral et al., 2008), with the former being more important in the southern sub-basins and the latter in the northern ones (Doyle and Barros, 2011) (see section 27.2.1). Increasing trends in streamflows have also been found in the Patos Lagoon in southern Brazil (Marques, 2012) and Laguna Mar Chiquita (a closed lake), and in the Santa Fe Province, both in Argentina, with ecological and erosive consequences (Bucher and Curto, 2012; Pasquini et al., 2006; Rodrigues Capítulo et al., 2010; Troin et al., 2010; Venencio and García, 2011). There is no clear long-term trend for the Amazon River. Espinoza et al. (2009a; 2011) showed that the 1974-2004 apparent stability in mean discharge at the main stem of the Amazon in Obidos is explained by opposing regional features of Andean rivers (e.g. increasing trends during the high-water period in Peruvian and Colombian Amazons and decreasing trend during the low-water period in Peruvian and Bolivian Amazons (Lavado et al., 2012). In recent years extremely low levels were experienced during the droughts of 2005 and 2010, while record high levels were detected during the 2009 and 2012 floods (see section 27.2.1). Major Colombian rivers draining to the Caribbean Sea (Magdalena and Cauca) exhibit decreasing trends along their main channels (Carmona and Poveda, 2011), while significant trends are absent for all other major large rivers in the Brazilian North East, and northern SA (Dai et al., 2009). Dai (2011) showed a drying trend in CA rivers. A rapid retreat and melting of the tropical Andes glaciers of Venezuela, Colombia, Ecuador, Peru and Bolivia has been further reported following the IPCC AR4, through use of diverse techniques (high confidence based on high agreement and robust evidence). Rabatel et al. (2013) provides a synthesis of these studies (specific papers are presented in Table 27-3a). Tropical glaciers retreat has accelerated in the second half of the 20th century (area loss between 20-50%), especially since the late 1970s in association with increasing temperature in the same period (Bradley et al., 2009). In early stages of glacier retreat associated streamflow tends to increase due to an acceleration of glacier melt, but after a peak in streamflow as the glacierized water reservoir gradually empties, runoff tends to decrease, as evidenced in the Cordillera Blanca of Peru (Baraer et al., 2012; Chevallier et al., 2011), where seven out of nine river basins have probably crossed a critical threshold, exhibiting a decreasing dry-season discharge Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (Baraer et al., 2012). Likewise, glaciers and icefields in the extra tropical Andes located in Central-South Chile and Argentina face significant reductions (see review in Masiokas et al. (2009) and details in Table 27-3b), with their effect being compounded by changes in snowpack extent, thus magnifying changes in hydrograph seasonality by reducing flows in dry seasons and increasing them in wet seasons (Pizarro et al., 2013; Vicuna et al., 2013). Central- South Chile and Argentina also face significant reductions in precipitation as shown in section 27.2.1, contributing to runoff reductions in the last decades of the 20th century (Rubio-Álvarez and McPhee, 2010; Seoane and López, 2007; Urrutia et al., 2011; Vicuna et al., 2013), corroborated with long-term trends found through dendrochronology (Lara et al., 2007; Urrutia et al., 2011). Trends in precipitation and runoff are less evident in the Central-North region in Chile (Fiebig-Wittmaack et al., 2012; Souvignet et al., 2012). [INSERT TABLE 27-3 HERE Table 27-3: Observed trends related to Andean cryosphere: a) Andean tropical glacier trends, and b) extra tropical Andean cryosphere (glaciers, snowpack, runoff effects) trends.] As presented in Table 27-4, the assessment of future climate scenarios implications in hydrologic related conditions shows a large range of uncertainty across the spectrum of climate models (mostly using CMIP3 simulations with the exception of Demaria et al. (2013) and scenarios considered. Nohara et al. (2006) studied climate change impacts on 24 of the main rivers in the world considering a large number of GCMs, and found no robust change for the Paraná (La Plata Basin) and Amazon Rivers. Nevertheless, in both cases the average change showed a positive value consistent at least with observations for the La Plata Basin. In a more recent work Nakaegawa et al. (2013a) showed statistically significant increase for both basins in a study that replicated that of Nohara et al. (2006) but with different hydrologic model. Focusing in extreme flows Guimberteau et al. (2013) show that by the middle of the century no change is found in high flow on the main stem of the Amazon River but there is a systematic reduction in low-flow streamflow. In contrast the northwestern part of the Amazon River shows a consistent increase in high flow and inundated area (Guimberteau et al., 2013; Langerwisch et al., 2013). On top of such climatic uncertainty, future streamflows and water availability projections are confounded by the potential effects of land use changes (Coe et al., 2009; Georgescu et al., 2013; Moore et al., 2007). [INSERT TABLE 27-4 HERE Table 27-4: Synthesis of projected climate change impacts on hydrologic related variables in CA and SA basins and major glaciers.] The CA region shows a consistent future runoff reduction. Maurer et al. (2009)studied climate change projections for the Lempa River basin, one of the largest basins in CA, covering portions of Guatemala, Honduras and El Salvador. They showed that future climate projections (increase in evaporation and reduction in precipitation) imply a reduction of 20% in inflows to major reservoirs in this system (see Table 27-4). Imbach et al. (2012) found similar results using a modeling approach that also considered potential changes in vegetation. These effects could have large hydropower generation implications as discussed in the case study (see section 27.6.1). The evolution of tropical Andes glaciers associated future climate scenarios has been studied using trend (e.g. Poveda and Pineda (2009), regression (e.g. Juen et al. (2007) and Chevallier et al. (2011) and explicit modeling (e.g. Condom et al., 2012) analysis. These studies indicate that glaciers will continue their retreat (Vuille et al., 2008a) and even disappear as glacier Equilibrium Line Altitudes rises, with larger hydrological effects during the dry season (Gascoin et al., 2011; Kaser et al., 2010). This is expected to happen during the next 20-50 years (Chevallier et al., 2011; Juen et al., 2007) (see Table 27-4). After that period water availability during the dry months is expected to diminish. A projection by Baraer et al. (2012) for the Santa River in the Peruvian Andes finds that once the glaciers are completely melt, annual discharge would decrease by 2% 30%, depending on the watershed. Glacier retreat can exacerbate current water resources related vulnerability (Bradley et al., 2006; Casassa et al., 2007; Mulligan et al., 2010; Vuille et al., 2008b), diminishing the mountains water regulation capacity, making it more expensive and less reliable the supply of water for diverse purposes, as well as for ecosystems integrity (Buytaert et al., 2011). Impacts on economic activities associated with conceptual scenarios of glacier melt reduction have been monetized (Vergara et al., 2007), representing about US$100 million in the case of water supply for Quito, and between US$212 million to US$ 1.5 billion in the case of the Peruvian electricity sector due to losses of hydropower generation (see case Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 study 27.6.1). Andean communities will face an important increase in their vulnerability, as documented by Mark et al. (2010), Pérez et al. (2010) and Buytaert and De Bievre (2012). In Central Chile, Vicuna et al. (2011) project changes in the seasonality of streamflows of the upper snowmelt- driven watersheds of the Limarí River, associated with temperature increases and reductions in water availability owing to a reduction (increase) in precipitation (evapotranspiration). Similar conclusions are derived across the Andes on the Limay River in Argentina by Seoane and López (2007). Under these conditions, semiarid highly populated basins (e.g. Santiago, Chile) and with extensive agriculture irrigation and hydropower demands are expected to increase their current vulnerability (ECLAC, 2009a; Fiebig-Wittmaack et al., 2012; Souvignet et al., 2010; Vicuna et al., 2012) (high confidence) (see Table 27-4). Projected changes in the cryosphere conditions of the Andes could affect the occurrence of extreme events, such as extreme low and high flows (Demaria et al., 2013). Glacial-lake outburst floods occurring in the icefields of Patagonia (Dussaillant et al., 2010; Marín et al., 2013), volcanic collapse and debris flow associated with accelerated glacial melting in the tropical Andes (Carey, 2005; Carey et al., 2012b; Fraser, 2012), and with volcanoes in southern Chile and Argentina (Tormey, 2010), as well as scenarios of water quality pollution by exposure to contaminants owing to glaciers retreat (Fortner et al., 2011). Another semiarid region that has been studied thoroughly is the NEBian (Hastenrath, 2012). De Mello et al. (2008), Gondim et al. (2008), Souza et al. (2010) and Montenegro and Ragab (2010) have shown that future climate change scenarios would decrease water availability for agriculture irrigation owing to reductions in precipitation and increases in evapotranspiration (medium confidence). Krol and Bronstert (2007) and Krol et al. (2006) presented an integrated modeling study that linked projected impacts on water availability for agriculture with economic impacts that could potentially drive full-scale migrations in the Brazilian northeast region. 27.3.1.2. Adaptation Practices At an institutional level, a series of policies have been developed to reduce vulnerability to climate variability as faced today in different regions and settings. In 1997, Brazil instituted the National Water Resources Policy and created the National Water Resources Management System under the shared responsibility between the States and the Federal government. Key to this new regulation has been the promotion of decentralization and social participation through the creation of National Council of Water Resources and their counterparts in the states, the States Water Resources Councils. The challenges and opportunities dealing with water resources management in Brazil in the face of climate variability and climate change have been well studied (Abers, 2007; Engle et al., 2011; Kumler and Lemos, 2008; Lorz et al., 2012; Medema et al., 2008). Other countries in the region are following similar approaches. In the last years, there have been constitutional and legal reforms towards more efficient and effective water resources management and coordination among relevant actors in Honduras, Nicaragua, Ecuador, Peru, Uruguay, Bolivia and Mexico; although in many cases, these innovations have not been completely implemented (Hantke Domas, 2011). Institutional and governance improvements are required to assure an effective implementation of these adaptation measures (e.g. Halsnaes and Verhagen, 2007; Engle and Lemos, 2010; Lemos et al., 2010; Zagonari, 2010; Pittock, 2011; and Kirchhoff et al. 2013). With regards to region specific freshwater resources issues it is important to consider adaptation to reduce vulnerabilities in the communities along the tropical Andes and the semi-arid basins in Chile-Argentina, North East Brazil, and the northern CA basins. Different issues have been addressed in the assessment of adaptation strategies for tropical Andean communities such as the role of governance and institutions (Lynch, 2012; Young and Lipton, 2006), technology (Carey et al., 2012a), and the dynamics of multiple stressors (Bury et al., 2013; McDowell and Hess, 2012). Semiarid regions are characterized by pronounced climatic variability and often by water scarcity and related social stress (Krol and Bronstert, 2007; Scott et al., 2012; 2013). Adaptation tools to face the threats of climate change for the most vulnerable communities in the Chilean semi-arid region are discussed by Young et al. (2010) and Debels et al. (2009). In CA, Benegas et al. (2009), Manuel-Navarrete et al. (2007) and Aguilar et al. (2009) provide different frameworks to understand vulnerability and adaptation strategies to climate change and variability in urban and rural contexts, although no specific adaptation strategies are suggested. The particular experience in NEB provides other examples of adaptation strategies to manage actual climate variability. Broad et al. (2007) and Sankarasubramanian et al. (2009) studied the potential benefits of streamflow forecast as a way to Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 reduce the impacts of climate change and climate variability on water distribution under stress conditions. An historical review and analysis of drought management in this region are provided by Campos and Carvalho (2008). Souza Filho and Brown (2009) studied different water distribution policy scenarios finding that the best option depended on the degree of water scarcity. The study by Nelson and Finan (2009) provides a critical perspective of drought-related policies, arguing that they constitute an example of maladaptation as they do not try to solve the causes of vulnerability and instead undermine resilience. Tompkins et al. (2008) are also critical of risk reduction practices in this region because they have fallen short of addressing the fundamental causes of vulnerability needed for efficient longer-term drought management. Other types of adaptation options that stem from studies on arid and semiarid regions are related to: a) increase in water supply from groundwater pumping (Burte et al., 2011; Döll, 2009; Kundzewicz and Döll, 2009; Nadal et al., 2013; Zagonari, 2010); fog interception practices (Holder, 2006; Klemm et al., 2012), and reservoirs and irrigation infrastructure (Fry et al., 2010; Vicuna et al., 2010; 2012); b) improvements in water demand management associated with increased irrigation efficiency and practices (Bell et al., 2011; Geerts et al., 2010; Jara-Rojas et al., 2012; Montenegro and Ragab, 2010; Van Oel et al., 2010), and changes towards less water intensive crops (Montenegro and Ragab, 2010). Finally flood management practices also provide a suite of options to deal with actual and future vulnerabilities related to hydrologic extremes, such as the management of ENSO-related events in Peru via participatory (Warner and Oré, 2006) or risk reduction approaches (Khalil et al., 2007), the role of land use management (Bathurst et al., 2010; 2011; Coe et al., 2011), and flood hazard assessment (Mosquera-Machado and Ahmad, 2006) (medium confidence). 27.3.2. Terrestrial and Inland Water Systems 27.3.2.1. Observed and Projected Impacts and Vulnerabilities CA and SA house the largest biological diversity and several of the world s megadiverse countries (Guevara and Laborde, 2008; Mittermeier et al., 1997). However, land use change has led to the existence of six biodiversity hotspots, i.e. places with a great species diversity that show high habitat loss and also high levels of species endemism: Mesoamerica, Chocó-Darien-Western Ecuador, Tropical Andes, Central Chile, Brazilian Atlantic forest, and Brazilian Cerrado (Mittermeier et al., 2005). Thus, conversion of natural ecosystems is the main proximate cause of biodiversity and ecosystem loss in the region (Ayoo, 2008). Tropical deforestation is the second largest driver of anthropogenic climate change on the planet, adding up to 17%-20% of total greenhouse gas emissions during the 1990s (Gullison et al., 2007; Strassburg et al., 2010). In parallel, the region has still large extensions of wilderness areas for which the Amazon is the most outstanding example. Nevertheless, some of these areas are precisely the new frontier of economic expansion. For instance, between 1996 and 2005 Brazil deforested about 19,500 km2 per year, which represented 2% to 5% of global annual CO2 emissions (Nepstad et al., 2009). Between 2005 and 2009, deforestation in the Brazilian Amazon dropped by 36%, which is partly related to the network of protected areas that now covers around 45.6% of the biome in Brazil (Soares-Filho et al., 2010). Using LandSHIFT modeling framework for land use change and the IMPACT projections of crop/livestock production, Lapola et al. (2011) projected that zero deforestation in the Brazilian Amazon forest by 2020 (and of the Cerrado by 2025) would require either a reduction of 26% 40% in livestock production until 2050 or a doubling of average livestock density from 0.74 to 1.46 head per hectare. Thus, climate change may imply reduction of yields and entail further deforestation. Local deforestation rates or rising greenhouse gases globally drive changes in the regional SA that during this century might lead the Amazon rainforest into crossing a critical threshold at which a relatively small perturbation can qualitatively alter the state or development of a system (Cox et al., 2000; Lenton et al., 2008; Nobre and Borma, 2009; Salazar et al., 2007; Sampaio et al., 2007). Various models are projecting a risk of reduced rainfall and higher temperatures and water stress, that may lead to an abrupt and irreversible replacement of Amazon forests by savanna-like vegetation, under a high emission scenario (A2), from 2050-2060 to 2100 (Betts et al., 2004; 2008; Cox et al., 2004; Malhi et al., 2008; Malhi et al., 2009; Marengo et al., 2011c; Nobre and Borma, 2009; Salazar et al., 2007; Sampaio et al., 2007; Sitch et al., 2008). The possible savannization or die-back of the Amazon region would potentially have large-scale impacts on climate, biodiversity and people in the region. The possibility of this Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 die-back scenario occurring, however, is still an open issue and the uncertainties are still very high (Rammig et al., 2010; Shiogama et al., 2011). Plant species are rapidly declining in CA, SA, Central and West Africa, and Southeast Asia (Bradshaw et al., 2009). Risk estimates of plant species extinction in the Amazon, which do not take into account possible climate change impacts, range from 5%-9% by 2050 with a habitat reduction of 12%-24% (Feeley and Silman, 2009) to 33% by 2030 (Hubbell et al., 2008). The highest percentage of rapidly declining amphibian species occurs in CA and SA. Brazil is among the countries with most threatened bird and mammal species (Bradshaw et al., 2009). A similar scenario is found in inland water systems. Among the components of aquatic biodiversity, fish are the best-known organisms (Abell et al., 2008) with Brazil accounting for the richest icthyofauna of the planet (Nogueira et al., 2010). For instance, the 540 Brazilian small microbasins host 819 fish species with restrict distribution. However, 29% of these microbasins have historically lost more than 70% of their natural vegetation cover and only 26% show a significant overlap with protected areas or indigenous reserves. Moreover, 40% of the microbasins overlap with hydrodams (see 27.6.1 and Chapter 3) or have few protected areas and high rates of habitat loss (Nogueira et al., 2010). The faster and more severe the rate of climate change, the more severe the biological consequences such as species decline (Brook et al., 2008). Vertebrate fauna in North and South America is projected to suffer species losses until 2100 of at least 10%, as forecasted in over 80% of the climate projections based on low emissions scenario (Lawler et al., 2009). Vertebrate species turnover until 2100 will be as high as 90% in specific areas of CA and the Andes Mountains for emission scenarios varying from low B1 to mid-high A2 (Lawler et al., 2009). Elevational specialists, i.e. a small proportion of species with small geographic ranges restricted to high mountains, are most frequent in the Americas (e.g. Andes and Sierra Madre) and might be particularly vulnerable to global warming because of their small geographic ranges and high energetic and area requirements, particularly birds and mammals (Laurance et al., 2011). In Brazil, projections for Atlantic forest birds (Anciaes and Peterson, 2006), endemic bird species (Marini et al., 2009), and plant species (by 2055, scenarios HHGSDX50 and HHGGAX50; Siqueira and Peterson, 2003) of the Cerrado indicate that distribution will dislocate towards the South and Southeast, precisely where fragmentation and habitat loss are worse. Global climate change is also predicted to increase negative impacts worldwide, including SA, on freshwater fisheries due to alterations in physiology and life histories of fish (Ficke et al., 2007). In addition to climate change impacts at individual species level, biotic interactions will be affected. Modifications in phenology, structure of ecological networks, predator-preys interactions and non-trophic interactions among organisms have been forecasted (Brooker et al., 2008; Walther, 2010). The outcome of non-trophic interactions among plants is expected to shift along with variation in climatic parameters, with more facilitative interactions in more stressful environments, and more competitive interactions in more benign environments (Anthelme et al., 2012; Brooker et al., 2008). These effects are expected to have a strong influence of community and ecosystem (re-) organization given the key engineering role played by plants on the functioning of ecosystems (Callaway, 2007). High Andean ecosystems, especially those within the tropics, are expected to face exceptionally strong warming effects during the 21th century because of their uncommonly high altitude (Bradley et al., 2006). At the same time they provide a series of crucial ecosystem services for millions people (Buytaert et al., 2011). For these reasons shifts in biotic interactions are expected to have negative consequences on biodiversity and ecosystem services in this region. Although in the region biodiversity conservation is largely confined to protected areas, with the magnitude of climatic changes projected for the century, it is expected that many species and vegetational types will lose representativeness inside such protected areas (Heller and Zavaleta, 2009). 27.3.2.2. Adaptation Practices The sub-set of practices that are multi-sectoral, multi-scale, and based on the premise that ecosystem services reduce the vulnerability of society to climate change are known as Ecosystem-based Adaptation (EbA) (Vignola et al., 2009; see also Glossary and CC-EA). Schemes such as the payment for environmental services (PES) and Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 community management fit the concept of EbA that begins to spread in CA and SA (Vignola et al., 2009). The principle behind these schemes is the valuation of ecosystem services that should reflect both the economic and cultural benefits derived from the human-ecosystem interaction and the capacity of ecosystems to secure the flow of these benefits in the future (Abson and Termansen, 2011). Since PES schemes have developed more commonly in CA and SA than in other parts of the world (Balvanera et al., 2012), this topic will be covered as a case study (see 27.6.2). Ecological restoration, conservation in protected areas, and community management can all be important tools for adaptation. A meta-analysis of 89 studies by Benayas et al. (2009) (with timescale of restoration varying from <5 to 300 years), including many in SA, showed that ecological restoration enhances the provision of biodiversity and environmental services by 44% and 25%, respectively, as compared to degraded systems (Benayas et al., 2009). Moreover, ecological restoration increases the potential for carbon sequestration and promotes community organization, economic activities and livelihoods in rural areas (Chazdon, 2008), as seen in examples of the Brazilian Atlantic Forest (Calmon et al., 2011; Rodrigues et al., 2011). In that sense, Locatelli et al. (2011) revised several ecosystem conservation and restoration initiatives in CA and SA that simultaneously help mitigate and adapt to climate change. Chazdon et al. (2009) also highlight the potential of restoration efforts to build ecological corridors (see Harvey et al., 2008, for example in Central America). The effective management of natural protected areas and the creation of new protected areas within national protected area systems and community management of natural areas are also efficient tools to adapt to climate change and to reconcile biodiversity conservation with socio-economic development (e.g., Bolivian Andes - Hoffmann et al., 2011; Panama Oestreicher et al. 2009 ). Porter-Bolland et al. (2012) compared protected areas with areas under community management in different parts of the tropical world, including CA and SA, and found that protected areas have higher deforestation rates than areas with community management. Similarly, Nelson and Chomitz (2011) found for the region that (i) protected areas of restricted use reduced fire substantially, but multi-use protected areas are even more effective; and that (ii) in indigenous reserves the incidence of forest fire was reduced by 16% as compared to non-protected areas. This contrasts with the findings of Miteva et al. (2012) that found protected areas more efficient in constraining deforestation than other schemes. Other good examples of adaptive community management in the continent include community forest concessions (e.g., Guatemala; Radachowsky et al., 2012), multiple-use management of forests (Guariguata et al., 2012 ; see also examples in Brazil Klimas et al., 2012, Soriano et al., 2012, and Bolívia Cronkleton et al., 2012); and local communities where research and monitoring protocols are in place to pay the communities for collecting primary scientific data (Luzar et al., 2011). 27.3.3. Coastal Systems and Low-Lying Areas 27.3.3.1. Observed and Projected Impacts and Vulnerabilities Climate change is altering coastal and marine ecosystems (Hoegh-Guldberg and Bruno, 2010). Coral reefs (Chapter 5, Cross-Chapter Box CC-CR), seagrass beds, mangroves, rocky reefs and shelves, and seamounts have few to no areas left in the world that remain unaffected by human influence (Halpern et al., 2008). Anthropogenic drivers associated with climate change decreased ocean productivity, altered food web dynamics, reduced abundance of habitat-forming species, shifting species distributions, and greater incidence of disease (Hoegh-Guldberg and Bruno, 2010). Coastal and marine impacts and vulnerability are often associated with collateral effects of climate change such as sea-level rise, ocean warming and ocean acidification (Cross-Chapter Box CC-OA). Overfishing, habitat pollution and destruction, and the invasion of species also negatively impact biodiversity and the delivery of ecosystem services (Guarderas et al., 2008; Halpern et al., 2008). Such negative impacts lead to losses that pose significant challenges and costs for societies, particularly in developing countries (Hoegh-Guldberg and Bruno, 2010). For instance, the Ocean Health Index (Halpern et al., 2012) that measures how healthy the coupling of the human-ocean system is for every coastal country (including parameters related to climate change), indicates that CA countries rank amongst the lowest values. For SA, Suriname stands out with one of the highest scores. Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Coastal states of Latin America and the Caribbean have a human population of more than 610 million, 3/4 of who live within 200 km of the coast (Guarderas et al., 2008). For instance, studying seven countries in the region (El Salvador, Nicaragua, Costa Rica, Panama, Colombia, Venezuela, Ecuador), Lacambra and Zahedi (2011) found that more than 30% of the population lives in coastal areas directly exposed to climatic events. Large coastal populations are related to the significant transformation marine ecosystems have been undergoing in the region. Fish stocks, places for recreation and tourism, and controls of pests and pathogens are all under pressure (Guarderas et al., 2008; Mora, 2008). Moreover, SLR varied from 2 to 7 mm/yr between 1950 and 2008 in CA and SA. The Western equatorial border, influenced by the ENSO phenomenon, shows a lower variation (of about 1 mm yr-1) and a range of variation under El Nino events of the same order of magnitude that the sustained past changes (Losada et al., 2013). The distribution of population is a crucial factor for inundation impact, with coastal areas being non- homogeneously impacted. A scenario of 1m SLR would affect some coastal populations in Brazil and the Caribbean islands (see Figure 27-6). (ECLAC, 2011a) [INSERT FIGURE 27-6 HERE Figure 27-6: Current and predicted coastal impacts and coastal dynamics in response to climate change.] Coastal impacts - based on trends observed and projections, the figure shows how potential impacts may be distributed in the region. Three cases: a) flooding: since flooding probability increases with increasing sea-level, one may expect a higher probability of flooding in locations showing >40% of change over the last 60 years in the 100- years total sea-level (excluding hurricanes). The figure also identifies urban areas where the highest increase in flooding level has been obtained.; b) beach erosion: it increases with potential sediment transport, thus locations where changes in potential sediment transport have increased over a certain threshold have a higher probability to be eroded; c) sea-ports and reliability of coastal structures: the figure shows locations where, in the case of having a protection structure in place, there is a reduction in the reliability of the structures due to the increase in the design wave height estimates (ECLAC, 2011a). Coastal dynamics - information based on historical time series that have been obtained by a combination of data reanalysis, available instrumental information and satellite information. Advanced statistical techniques have been used for obtaining trends including uncertainties (Izaguirre et al., 2013; Losada et al., 2013).] The greatest flooding levels (hurricanes not considered) in the region are found in Rio de La Plata area, which combine a 5 mm yr-1 change in storm surge with SLR changes in extreme flooding levels (ECLAC, 2011a; Losada et al., 2013). Extreme flooding events may become more frequent since return periods are decreasing, and urban coastal areas in the eastern coast will be particularly affected, while at the same time beach erosion is expected to increase in southern Brazil and in scattered areas at the Pacific coast. (ECLAC, 2011a) The majority of literature concerning climate change impacts for coastal and marine ecosystems considers coral reefs (see also Chapter 5, Cross-Chapter Box CC-CR), mangroves and fisheries. Coral reefs are particularly sensitive to climate-induced changes in the physical environment (Baker et al., 2008) to an extent that 1/3 of the more than 700 species of reef-building corals worldwide are already threatened with extinction (Carpenter et al., 2008). Coral bleaching and mortality are often associated with ocean warming and acidification (Baker et al., 2008). If extreme sea surface temperatures are to continue, the projections of scenarios SRES (A1FI, 3C sensitivity, and A1B with 2C and 4.5C sensitivity) indicate that it is possible that the Mesoamerican coral reef will collapse by mid-century (between 2050 and 2070), causing major economic losses (WB, 2009). Extreme high sea surface temperatures have been increasingly documented in the western Caribbean near the coast of CA and have resulted in frequent bleaching events (1993, 1998, 2005, and again in 2010) of the Mesoamerican coral reef, located along the coasts of Belize, Honduras and Guatemala (Eakin et al., 2010). Reef but also mangrove ecosystems are estimated to contribute greatly to goods and services in economic terms. In Belize, for example, this amount is approximately US$395-US$559 million annually, primarily through marine-based tourism, fisheries and coastal protection (Cooper et al., 2008). In the Eastern Tropical Pacific, seascape trace abundance of cement and elevated nutrients in upwelled waters are factors that help explain high bioerosion rates of local coral reefs (Manzello et al., 2008). In the southwestern Atlantic coast, eastern Brazilian reefs might suffer a massive coral cover decline in the next 50 years Francini-Filho et al. (2008). This estimate is based on coral disease prevalence and progression rate, along with growth rate of Mussismilia braziliensis - a major reef-building coral species that is endemic in Brazil. These authors also pointed out that coral diseases intensified between 2005 and 2007 based on qualitative observations since the Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 1980s and regular monitoring since 2001. They have also predicted that - the studied coral species will be nearly extinct in less than a century if the current rate of mortality due to disease is not reversed. Mangroves are largely affected by anthropogenic activities whether or not they are climate driven. All mangrove forests, along with important ecosystem goods and services, could be lost in the next 100 years if the present rate of loss continues (1-2% a year) (Duke et al., 2007). Moreover, estimates are that climate change may lead to a maximum global loss of 10 15% of mangrove forest by 2100 (Alongi, 2008). In CA and SA, some of the main drivers of loss are deforestation and land conversion, agriculture and shrimp ponds (Polidoro et al., 2010). The Atlantic and Pacific coasts of CA are some of the most endangered in the planet with regards to mangroves, since approximately 40% of the present mangroves species are threatened with extinction (Polidoro et al., 2010). Approximately 75% of the mangrove extension of the planet is concentrated in 15 countries, among which Brazil is included (Giri et al., 2011). The rate of survival of original mangroves lies between 12.8% and 47.6% in the Tumaco Bay (Colombia), resulting in ecosystem collapse, fisheries reduction and impacts on livelihoods (Lampis, 2010). Gratiot et al. (2008) project for the current decade an increase of mean high water levels of 6 cm followed by 90m shoreline retreat implying flooding of thousands of hectares of mangrove forest along the coast of French Guiana. Peru and Colombia are two of the eight most vulnerable countries to climate change impacts on fisheries, due to the combined effect of observed and projected warming, to species and productivity shifts in upwelling systems, to the relative importance of fisheries to national economies and diets, and limited societal capacity to adapt to potential impacts and opportunities (Allison et al., 2009). Fisheries production systems are already pressured by overfishing, habitat loss, pollution, invasive species, water abstraction and damming (Allison et al., 2009). In Brazil, a decadal rate of 0.16 trophic level decline (as measured by the Marine Trophic Index, which refers to the mean trophic level of the catch) has been detected through most of the northeastern coast, between 1978 and 2000, which is one of the highest rates documented in the world (Freire and Pauly, 2010). Despite the focus in the literature on corals, mangroves and fisheries, there is evidence that other benthic marine invertebrates that provide key services to reef systems, such as nutrient cycling, water quality regulation, and herbivory, are also threatened by climate change (Przeslawski et al., 2008). The same applies for seagrasses for which a worldwide decline has accelerated from a median of 0.9% yr-1 before 1940 to 7% yr-1 since 1990, which is comparable to rates reported for mangroves, coral reefs, tropical rainforests and place seagrass meadows among the most threatened ecosystems on earth (Waycott et al., 2009). A major challenge of particular relevance at local and global scales will be to understand how these physical changes will impact the biological environment of the ocean (e.g., Gutiérrez et al., 2011b), as the Humboldt Current system -flowing along the west coast of SA- is the most productive upwelling system of the world in terms of fish productivity. 27.3.3.2. Adaptation Practices Designing marine protected areas (MPAs) that are resilient to climate change is a key adaptation strategy in coastal and marine environments (McLeod et al., 2009). By 2007, Latin America and the Caribbean (which includes CA and SA countries) had over 700 MPAs established covering around 1.5% of the coastal and shelf waters, most of which allow varying levels of extractive activities (Guarderas et al., 2008). This protected area cover, however, is insufficient to preserve important habitats or connectivity among populations at large biogeographic scales (Guarderas et al., 2008). Nevertheless, examples of adaptation in CA and SA are predominantly related to MPAs. In Brazil, a protected area type known as Marine Extractive Reserves currently benefits 60,000 small-scale fishermen along the coast (Moura et al., 2009). Examples of fisheries co-management, a form of a participatory process involving local fishermen communities, government, academia and NGOs, are reported to favor a balance between conservation of marine fisheries, coral reefs and mangroves on the one hand (Francini-Filho and Moura, 2008), and the improvement of livelihoods, as well as the cultural survival of traditional populations on the other (Hastings, 2011; Moura et al., 2009). Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Significant financial and human resources are expended annually in the marine reserves to support reef management efforts. These actions, including the creation of marine reserves to protect from overfishing, improvement of watershed management, and protection or replanting of coastal mangroves, are proven tools to improve ecosystem functioning. In Mesoamerican reefs, (Carilli et al., 2009) found out that such actions may also actually increase the thermal tolerance of corals to bleaching stress and thus the associated likelihood of surviving future warming. In relation to mangroves, in addition to marine protected areas that include mangroves and functionally linked ecosystems, Gilman et al. (2008) list a number of other relevant adaptation practices: coastal planning to facilitate mangrove migration with sea-level rise, management of activities within the catchment that affect long-term trends in the mangrove sediment elevation, better management of non-climate stressors, and the rehabilitation of degraded areas. However, such types of practices are not frequent in the region. On the other hand, the implementation of adaptation strategies to sea level rise or to address coastal erosion is more commonly seen in many countries in the region (Lacambra and Zahedi, 2011). For instance, redirecting new settlements to better-protected locations and to promote investments in appropriate infrastructure shall be required in the low elevation coastal zones (LECZ) of the region, particularly in lower income countries with limited resources, which are especially vulnerable. The same applies to countries with high shares of land (e.g., Brazil ranking 7th worldwide of the total land area in the LECZ) and/or population (e.g., Guyana and Suriname rank 2nd and 5th by the share of population in the LECZ, having respectively 76% and 55% of their populations in such areas (McGranahan et al., 2007). Adaptation will demand effective and enforceable regulations and economic incentives, all of which require political will as well as financial and human capital (McGranahan et al., 2007). Adaptive practices addressing river flooding are also being made available as in the study of Casco et al. (2011) for the low Parana river, in Argentina; (see also Chapters 5 and 6 for coastal and marine adaptation). 27.3.4. Food Production Systems and Food Security 27.3.4.1. Observed and Projected Impacts and Vulnerabilities Increases in the global demand for food and biofuels promoted a sharp increase in agricultural production in SA and CA mainly associated with the expansion of planted areas (see Chapter 7), and this trend is predicted to continue in the future (see 27.2.2.1). Ecosystems are being and will be affected in isolation and synergistically by climate variability/change and land use changes, which are comparable drivers of environmental change (see 27.2.2.1; 27.3.2.1). By the end of 21th century (13 GCMs, under SRES A1B and B1) SA could lose between 1% and 21% of its arable land due to CC and population growth (Zhang and Cai, 2011). Optimal land management could combine efficient agricultural and biofuels production with ecosystem preservation under CC, however current practices are leading to a deterioration of ecosystems throughout the continent (see section 27.3.2). In southern Brazilian Amazonia water yields (mean daily discharge (mm.day-1)) were near four times higher in soy than forested watersheds, and showed greater seasonal variability (Hayhoe et al., 2011). In the Argentinean Pampas current land use changes disrupt water and biogeochemical cycles and may result in soil salinization, altered C and N storage, surface runoff and stream acidification (Berthrong et al., 2009; Farley et al., 2009; Nosetto et al., 2008). In central Argentina flood extension was associated with the dynamics of groundwater level that has been influenced by precipitation and land use change (Viglizzo et al., 2009). Observed impacts: The SESA region has shown significant increases in precipitation and wetter soil conditions during the 20th century (Giorgi, 2002) (see Table 27-1) that benefited summer crops and pastures productivity, and contributed to the expansion of agricultural areas (Barros, 2010; Hoyos et al., 2012). Wetter conditions observed during 1970-2000 (in relation to 1930-1960) led to increases in maize and soybean yields (9% to 58%) in Argentina, Uruguay and Southern Brazil (Magrin et al., 2007b). Even if rainfall projections estimate increases of about 25% in SESA for 2100, agricultural systems could be threatened if climate reverts to a drier situation due to interdecadal variability. This could put at risk the viability of continuous agriculture in marginal regions of the Argentina s Pampas (Podestá et al., 2009). During the 30 s and 40 s, dry and windy conditions together with deforestation, Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 overgrazing, overcropping, and non-suitable tillage produced severe dust storms, cattle mortality, crop failure, and rural migration (Viglizzo and Frank, 2006). At the global scale (see Chapter 7), warming since 1981 has reduced wheat, maize and barley productivity, although the impacts were small compared with the technological yield gains over the same period (Lobell and Field, 2007). In central Argentina, simulated potential wheat yield without considering technological improvements- has been decreasing at increasing rates since 1930 (1930-2000: -28 kg/ha/year; 1970-2000: -53 kg/ha/year) in response to increases in minimum temperature during October-November (1930-2000: +0.4°C/decade; 1970-2000: +0.6°C/decade) (Magrin et al., 2009). The observed changes in the growing season temperature and precipitation between 1980 and 2008 have slowed the positive yield trends due to improved genetics in Brazilian wheat, maize and soy, as well as Paraguayan soy. In contrast, rice in Brazil and soybean in Argentina have benefited from precipitation and temperature trends (2011). In Argentina, increases in soybean yield may be associated with weather types that favour the entry of cold air from the south reducing thermal stress during flowering and pod set, and weather types that increase the probability of dry days at harvest (Bettolli et al., 2009). Projected impacts: The assessment of future climate scenarios implications in food production and food security (see Table 27-5) shows a large range of uncertainty across the spectrum of climate models and scenarios. One of the uncertainties is related to the effect of CO2 on plant physiology. Many crops (such as soybean, common bean, maize and sugarcane) can probably respond with an increasing productivity as a result of higher growth rates and better water use efficiency. However, food quality could decrease due to higher sugar contents in grain and fruits, and decreases in the protein content in cereals and legumes (DaMatta et al., 2010). Uncertainties associated with climate and crop models, as well as with the uncertainty in human behavior, potentially lead to large error bars on any long- term prediction of food output. However, the trends presented here represent the current available information. In SESA, some crops could be benefited until mid-21st-century if CO2 effects are considered (see Table 27-5), although interannual and decadal climate variability could provoke important damages. In Uruguay and Argentina productivity could increase or remain almost stable until the 2030s-2050s depending on the SRES scenario (ECLAC, 2010c). Warmer and wetter conditions may benefit crops towards the southern and western zone of the Pampas (ECLAC, 2010c; Magrin et al., 2007c). In South Brazil, irrigated rice yield (Walter et al., 2010) and bean productivity (Costa et al., 2009) is expected to increase. If technological improvement is considered, the productivity of common bean and maize could increase between 40% and 90% (Costa et al., 2009). Sugarcane production could benefit as warming could allow the expansion of planted areas towards the south, where low temperatures are a limiting factor (Pinto et al., 2008). Increases in crop productivity could reach 6% in Sao Paulo state towards 2040 (Marin et al., 2009). In Paraguay the yields of soybean, maize and wheat could have slight variations (-1.4% to +3.5%) until 2020 (ECLAC, 2010a). In Chile and western Argentina, yields could be reduced by water limitation. In central Chile (30S to 42S) temperature increases, reduction in chilling hours and water shortages may reduce productivity of winter crops, fruits, vines and radiata pine. Conversely, rising temperatures, more moderate frosts and more abundant water will very likely benefit all species towards the South (ECLAC, 2010a; Meza and Silva, 2009). In northern Patagonia (Argentina) fruit and vegetable growing could be negatively affected because of a reduction in rainfall and in average flows in the Neuquén River basin. In the north of the Mendoza basin (Argentina) increases in water demand, due to population growth, may compromise the availability of subterranean water for irrigation, pushing up irrigation costs and forcing many producers out of farming towards 2030. Also, water quality could be reduced by the worsening of existing salinization processes (ECLAC, 2010a). In CA, NEB and parts of the Andean region (Table 27-5) CC could affect crop yields, local economies and food security. It is very likely that growing season temperatures in parts of tropical SA, east of the Andes and CA exceed the extreme seasonal temperatures documented from 1900 to 2006 at the end of this century (23 GCMs), affecting regional agricultural productivity and human welfare (Battisti and Naylor, 2009). For NEB, declining crop yields in subsistence crops such as beans, corn and cassava are projected (Lobell et al., 2008; Margulis et al., 2010). In addition, increases in temperature could reduce the areas currently favorable to cowpea bean (Silva et al., 2010). The highest warming foreseen for 2100 (5.8 °C, under SRES A2 scenario) could make the coffee crop unfeasible in Minas Gerais and Sao Paulo (SE Brazil) if no adaptation action is accomplished. Thus, the coffee crop may have to Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 be transferred to southern regions where temperatures are lower and the frost risk will be reduced (Camargo, 2010). With +3C, Arabica coffee is expected to expand in the extreme south of Brazil, the Uruguayan border and North of Argentina (Zullo et al., 2011). Brazilian potato production could be restricted to a few months in currently warm areas, which today allow potato production all around the year (Lopes et al., 2011). Large losses of suitable environments for the Pequi tree (Caryocar brasiliense; an economically important Cerrado fruit tree) are projected by 2050, mainly affecting the poorest communities in Central Brazil (Nabout et al., 2011). In the Amazon region soybean yields would be reduced by 44% in the worst scenario (HadCM3 and no CO2 fertilization) by 2050 (Lapola et al. (2011). By 2050, according to 17GCMs under SRES A2 scenario, 80% of crops will be impacted in more than 60% of current areas of cultivation in Colombia, with severe impacts in perennial and exportable crops (Ramirez- Villegas et al., 2012). Teixeira et al. (2013) identified hot spots for heat stress towards 2071-2100 under the A1B scenario and suggest that rice in South East Brazil, maize in CA and SA, and soybean in Central Brazil will be the crops and zones most affected by increases in temperature. In CA, changes projected in climate could severely affect the poorest population and specially their food security increasing the current rate of chronic malnutrition. Currently, Guatemala is the most food insecure country by percentage of the population (30.4%) and the problem has been increasing in recent years (FAO, 2012). The impact of climate variability and change is a great challenge in the region. As an example, the recent rust problem on the coffee sector of 2012/2013 has affected near to 600.000 ha (55% of the total area) (ICO, 2013) and will reduce employment by 30% to 40% for the harvest 2013/2014 (FEWS NET, 2013). At least 1.4 million people in Guatemala, El Salvador, Honduras and Nicaragua depend on the coffee sector, which is very susceptible to climate variations. In Panamá, the large interannual climate variability will continue to be the dominant in uence on seasonal maize yield into the coming decades (Ruane et al., 2013). In the future, warming conditions combined with more variable rainfall are expected to reduce maize, bean and rice productivity (ECLAC, 2010c); rice and wheat yields could decrease up to 10% by 2030 (Lobell et al., 2008 (medium confidence) (see Table 26-7). In CA, near to 90% of agricultural production destined to internal consumption is composed by maize (70%), bean (25%) and rice (6%) (ECLAC, 2011d). [INSERT TABLE 27-5 HERE Table 27-5: Impacts on agriculture.] CC may also alter the current scenario of plant diseases and their management, having effects on productivity (Ghini et al., 2011). In Argentina, years with severe infection of late cycle diseases in soybean could increase; severe outbreaks of the Mal de Rio Cuarto virus in maize (natural vectors: Delphacodes kuscheli and Delphacodes hayward) could be more frequent; and wheat head fusariosis will increase slightly in the south of the Pampas region by the end of the century (ECLAC, 2010a). In Brazil favorable areas for soybean and coffee rusts will move toward the south, particularly for the hottest scenario of 2080 (Alves et al., 2011). Potato late blight (Phytophtora infestans) severity is expected to increase in Perú (Giraldo et al., 2010). The choice of livestock species could change in the future. For example, by 2060, under a hot and dry scenario, beef and dairy cattle, pigs and chickens production choice could decrease between 0.9 and 3.2%, while sheep election could increase by 7% mainly in the Andean countries (Seo et al., 2010). Future climate could strongly affect milk production and feed intake in dairy cattle in Brazil, where substantial modifications in areas suitable for livestock, mainly in the Pernambuco region, are expected (Silva et al., 2009). Warming and drying conditions in Nicaragua could reduce milk production, mainly among farmers that are already seriously affected under average dry season conditions (Lentes et al., 2010). CC impact on regional welfare will depend not only on changes in yield, but also in international trade. According to Hertel et al. (2010), by 2030, global cereal price could change between increases of 32% (low-productivity scenario) or decreases of 16% (optimistic yield scenario). A rise in prices could benefit net exporting countries like Brazil, where gains from terms of trade shifts could outweigh the losses due to CC. Despite experiencing significant negative yield shocks, some countries tend to gain from higher commodity prices. However, most poor household Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 are food purchasers and rising commodity prices tend to have a negative effect on poverty (von Braun, 2007). According to Chapter 7, increases in prices during 2007-2009 led to rising poverty in Nicaragua. 27.3.4.2. Adaptation Practices Genetic advances and suitable soil and technological management may induce an increase in some crops yield despite unfavorable future climate conditions. In Argentina, genetic techniques, specific scientific knowledge and land-use planning are viewed as promising sources of adaptation (Urcola et al., 2010). Adjustments in sowing dates and fertilization rates could reduce negative impacts or increase yields in maize and wheat crops in Argentina and Chile (Magrin et al., 2009; Meza and Silva, 2009; Travasso et al., 2009b). Furthermore, in central Chile and southern Pampas in Argentina warmer climates could allow performing two crops per season increasing productivity per unit land (Meza et al., 2008; Monzon et al., 2007). In Brazil, adaptation strategies for coffee crops include: planting at high densities, vegetated soil, accurate irrigation and breeding programs, and shading management system (arborization) (Camargo, 2010). Shading is also used in Costa Rica and Colombia. In South Brazil, a good option for irrigated rice could be to plant early cultivars (Walter et al., 2010). Water management is other option for a needed better preparedness regarding water scarcity (see section 27.3.1). In Chile, the adoption of water conservation practices depends on social capital, farm size and land use; and the adoption of technologies that require investment depend on the accesses to credit and irrigation water subsidies (Jara-Rojas et al., 2012). Deficit irrigation could be an effective measure for water savings in dry areas such as the Bolivian Altiplano (quinoa), central Brazil (tomatoes) and northern Argentina (cotton) (Geerts and Raes, 2009). In rainfed crops adaptive strategies might need to look at the harvest, storage, temporal transfer and efficient use of rainfall water. In addition, some agronomic practices like: fallowing, crop sequences, groundwater management, no- till operations, cover crops and fertilization could improve the adaptation to water scarcity (Quiroga and Gaggioli, 2011). One approach to adapting to future CC is by assisting people to cope with current climate variability (Baethgen, 2010), for which the use of climatic forecasts in agricultural planning presents a measure. Increased access and improvement of climate forecast information enhances the ability of the farmers in the Brazilian Amazon to cope with El Nino impacts (Moran et al., 2006). The Southern Oscillation Index for maize, and the South Atlantic Sea Surface Temperature for soybean and sunflower were the best indicators of annual crop yield variability in Argentina (Travasso et al., 2009a). Another possibility to cope with extreme events consists in transferring weather- related risks by using different types of rural insurance (Baethgen, 2010). Index insurance is one mechanism that has been recently introduced to overcome obstacles to traditional agricultural and disaster insurance markets (see chapter 15). For the support of such parametric agricultural insurance, a Central American climate database was recently established (SICA, 2013). Local and indigenous knowledge have the potential to bring solutions even in the face of rapidly changing climatic conditions (Altieri and Koohafkan, 2008; Folke et al., 2002); although, migration, climate change, and market integration are reducing indigenous capacity for dealing with weather and climate risk (Pérez et al., 2010; Valdivia et al., 2010). Crop diversification is used in the Peruvian Andes to suppress pest outbreaks and dampen pathogen transmission (Lin, 2011). In Honduras, Nicaragua and Guatemala traditional practices have proven more resilient to erosion and runoff and have helped retain more topsoil and moisture (Holt-Gimenez, 2002). In El Salvador, if local sustainability efforts continue the future climate vulnerability index could only slightly increase by 2015 (Aguilar et al., 2009). Studies with indigenous farmers in highland Bolivia and Peru indicate that constraints on access to key resources must be addressed for reducing vulnerability over time (McDowell and Hess, 2012; Sietz et al., 2012). In Guatemala and Honduras adaptive response between coffees farmers is mainly related to land availability, while participation in organized groups and access to information contribute to adaptive decision-making (Tucker et al., 2010). Otherwise, adaptation may include an orientation towards non-farming activities to sustain their livelihoods and be able to meet their food requirements (Sietz, 2011). In NEB increasing vulnerability related to degradation of natural resources (due to over use of soil and water) encouraged farmers toward off-farm activities, however they could not improve their well-being (Sietz et al., 2006; Sietz et al., 2011). Migration is other strategy in ecosystems and regions at high risk of climate hazards (see Section 27.3.1.1). During 1970-2000 LAC has had the great rate of Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 net migration per population in the dryland zones (Sherbinin et al., 2012). In CA near to 25% of the surveyed households reported some type of migration during the coffee crisis (Tucker et al., 2010). Some migrations, e.g. Guatemala 1960s-1990s; El Salvador 1950s-1980s; NEB 1960s-present, have provoked conflict in receiving areas (Reuveny, 2007). Shifting in agricultural zoning has been an autonomous adaptation observed in SA. In Argentina e.g., increases in precipitation promoted the expansion of the agricultural frontier to the West and North of the traditional agricultural area, resulting in environmental damage that could be aggravated in the future (Barros, 2008; República Argentina, 2007). Adjustment of production practices, like farmers in the semi-arid zones of mountain regions of Bolivia have begun as they noticed strong changes in the climate since the 1980s, including upward migration of crops, selection of more resistant varieties and water capturing, presents a further adaptation measure (PNCC, 2007). Organic systems could enhance adaptive capacity due to the application of traditional skills and farmers knowledge, soil fertility-building techniques and a high degree of diversity (ITC, 2007). As mentioned previously, crop diversity, local knowledge, soil conservation, and economic diversity are all documented strategies for managing risk in CA and SA. A controversial, but important issue in relation to adaptation is the use of genetically modified plants to produce food, with biotech crops being a strategy to cope with the needed food productivity increase considering global population trend (see Chapter 7). Brazil and Argentina are the 2nd and 3rd fastest growing biotech crop producers in the world after the US (Marshall, 2012). However, this option is problematic for the small farms (Mercer et al., 2012), which are least favorable towards GMO (??) (Soleri et al., 2008). According to Eakin and Wehbe (2009) some practices could be an adaptive option for specific farm enterprises, but may have maladaptive implications at regional scales, and over time, become maladaptive for individual enterprises. 27.3.5. Human Settlements, Industry, and Infrastructure According to the World Bank database (WB, 2012) CA and SA are the geographic regions with the second highest urban population (79%), behind North America (82%) and well above the world average (50%). Therefore this section focuses on assessing the literature on climate change impacts and vulnerability of urban human settlements. The information provided should be complemented with other sections of the chapter (see 27.2.2.2; 27.3.1; 27.3.3; and 27.3.7) 27.3.5.1. Observed and Projected Impacts and Vulnerabilities Urban human settlements suffer from many of the vulnerabilities and impacts already presented in several sections of this chapter. The provision of critical resources and services as already discussed in the chapter water, health and energy and of adequate infrastructure and housing remain determinants of urban vulnerability that are enhanced by climate change (Roberts, 2009; Romero-Lankao et al., 2013b; Romero-Lankao et al., 2012b; Smolka and Larangeira, 2008; Winchester, 2008). Water resource management for example (see section 27.3.1), is a major concern for many cities that need to provide both drinking water and sanitation (Henríquez Ruiz, 2009). More than 20% of the population in the region are concentrated in the largest city in each country (WB, 2012), hence water availability for human consumption in the region s megacities (e.g. Sao Paulo, Santiago, Lima, Buenos Aires) is of great concern. In this context, reduction in glacier and snowmelt related runoff in the Andes poses important adaptation challenges for many cities, e.g. the metropolitan areas of Lima, La Paz/El Alto and Santiago de Chile (Bradley et al., 2006; Hegglin and Huggel, 2008; Melo et al., 2010). Flooding is also a preoccupation in several cities. In Sao Paulo for example, according to Marengo et al. (2009b; 2013b) the number of days with rainfall above 50 mm were almost zero during the 1950s and now they occur between 2 to 5 times per year (2000-2010). The increase in precipitation is one of the expected risks affecting the city of Sao Paulo as presented in Box 27-2. Increases in flood events during 1980-2000 have been observed also in the Buenos Aires province and Metropolitan Area (Andrade and Scarpati, 2007; Barros et al., 2008; Hegglin and Huggel, 2008; Nabel et al., 2008). There are also the combined effects of climate change impacts, human settlements features and other stresses, such as more intense pollution events (Moreno, 2006; Nobre et al., Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2011; Nobre, 2011; Romero-Lankao et al., 2013b) and more intense hydrological cycles from urban heat-island effects. In terms of these combined effects, peri-urban areas and irregular settlements pose particular challenges to urban governance and risk management given their scale, lack of infrastructure and socio-economic fragility (Romero-Lankao et al., 2012). ______START BOX 27-2 HERE______ Box 27-2. Vulnerability of South American Megacities to Climate Change: The Case of the Metropolitan Region of Sao Paulo (MRSP) Research in the Metropolitan Region of Sao Paulo (MRSP), between 2009 and 2011, represents a comprehensive and interdisciplinary project on the impacts of climate variability and change, and vulnerability of Brazilian megacities. Studies derived from this project (Marengo et al., 2013b; Nobre et al., 2011) identify the impacts of climate extremes on the occurrence of natural disasters and human health. These impacts are linked to a projected increase of 38% in the extension of the urban area of the MRSP by 2030, accompanied by a projected increase in rainfall extremes. These may induce an intensification of urban flash floods and landslides, affecting large populated areas already vulnerable to climate extremes and variability. The urbanization process in the MRSP has been affecting the local climate, and the intensification of the heat island effect to a certain degree may be responsible for the 2°C warming detected in the city during the last 50 years (Nobre et al., 2011). This warming has been further accompanied by an increase in heavy precipitation as well as more frequent warm nights (Marengo et al., 2013b; Silva Dias et al., 2012). By 2100, climate projections based on data from 1933-2010 show an expected warming between 2-3°C in the MRSP, together with a possible doubling of the number of days with heavy precipitation in comparison to the present (Marengo et al., 2013b; Silva Dias et al., 2012). With the projected changes in climate and in the extension of the MRSP (Marengo et al., 2013b) more than 20% of the total area of the city could be potentially affected by natural disasters. More frequent floods may increase the risk of leptospirosis, which, together with increasing air pollution and worsening environmental conditions that trigger the risk of respiratory diseases, would leave the population of the MRSP more vulnerable. Potential adaptation measures include a set of strategies that need to be developed by the MRSP and its institutions to face these environmental changes. These include improved building controls to avoid construction in risk areas, investment in public transportation, protection of the urban basins and the creation of forest corridors in the collecting basins and slope regions. The lessons learned suggest that the knowledge on the observed and projected environmental changes, as well as on the vulnerability of populations living in risk areas is of great importance for defining adaptation policies that in turn constitute a first step towards building resilient cities that in turn improve urban quality of life in Brazil. ______END BOX 27-2 HERE______ Changes in prevailing urban climates have led to changing patterns of disease vectors, and water-borne disease issues linked to water availability and subsequent quality (see section 27.3.7). The influence of climate change on particulate matter and other local contaminants is another concern (Moreno, 2006; Romero-Lankao et al., 2013b). It is important to highlight the relationship between water and health, given the problems of water stress and intense precipitation events affecting many urban centers. Both relate to changing disease risks, as well as wider problems of event-related mortalities and morbidity, and infrastructure and property damage. These risks are compounded for low-income groups in settlements with little or no service provision, e.g. waste collection, piped drinking water, sanitation, (ECLAC, 2008). Existing cases of flooding, air pollution and heat waves reveal that not only low-income groups are at risk, but also that wealthier sectors are not spared. Factors such as high-density settlement (Barros et al., 2008) and the characteristics of some hazards explain this e.g., poor and wealthy alike are at risk from air pollution and temperature in Santiago de Chile and Bogota (2013b; Romero-Lankao et al., 2012a). There are also other climate change risks in terms of economic activity location and impacts on urban manufacturing and service workers, e.g. thermal stress (Hsiang, 2010), and the forms of urban expansion or sprawl into areas where ecosystem services may be compromised and risks enhanced, e.g. floodplains. Both processes are also related to rising motorization rates that facilitate suburban development and new regional agglomerations that bring pressure Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 to bear on land uses that favor infiltration, surface cooling and biodiversity; the number of light vehicles in Latin America and the Caribbean is expected to double between 2000 and 2030, and be three times the 2000 figure by 2050 (ECLAC, 2009c). While urban populations face diverse social, political, economic and environmental risks in daily life, climate change adds a new dimension to these risk settings (Pielke Jr et al., 2003; Roberts, 2009; Romero-Lankao and Qin, 2011). Since urban development remains fragile in many cases, with weak planning responses, climate change can compound existing challenges. The probabilities and magnitudes of these events in each urban center will differ significantly according to socioeconomic, institutional and physical contexts. 27.3.5.2. Adaptation Practices The direct and indirect effects of climate change as flooding, heat islands and food insecurity present cities with a set of challenges and opportunities for mainstreaming flood management, warning systems and other adaptation responses with sustainability goals (Bradley et al., 2006; Hardoy and Pandiella, 2009; Hegglin and Huggel, 2008; Romero-Lankao, 2010; Romero-Lankao et al., 2013a; Romero-Lankao, 2012). Urban populations, economic activities and authorities have a long experience of responding to climate-related hazards, particularly through disaster risk management (e.g., Tucuman and San Martin, Argentina (Plaza and Pasculi, 2007; Sayago et al., 2010)), and land use and economic development planning to a limited extent (Barton, 2009). Climate policies can build on these. Local administrations participate in the ICLEI, C40, IDB Emerging and Sustainable Cities Initiative (ESCI) (IDB, 2013), and other networks, demonstrating their engagement in the generation of more climate-resilient cities. In smaller settlements, there is less capacity for adequate responses (e.g., climate change and vulnerability information (Hardoy and Romero-Lankao, 2011). Policies, plans and programs are required to reduce social vulnerability, and identify and reduce potential economic effects of climate on the local economy. Rio de Janeiro, for example, with its coastline property and high dependence on tourists (and their perceptions of risk), cannot ignore these climate-related hazards (Gasper et al., 2011). Poverty and vulnerability, as interlinked elements of the adaptation challenge in CA and SA, remain pivotal to understanding how urban climate policies can be streamlined with broader development issues and not solely the capacity to respond to climate change (Hardoy and Pandiella, 2009; Hardoy and Romero-Lankao, 2011; Winchester and Szalachman, 2009). These broader links include addressing the determinants of vulnerability (e.g., access to education, health and infrastructure, and to emergency response systems (Romero-Lankao, 2007a; Romero-Lankao and Qin, 2011). Among these response options, a focus on social assets has been highlighted by Rubin and Rossing (2012), rather than a, purely, physical asset focus. Much urbanisation involves in-migrating or already resident, low-income groups and their location in risk-prone zones (Costa Fereira et al., 2011). The need to consider land use arrangements, particularly urban growth on risk- prone zones, as part of climate change adaptation highlights the role of green areas that mitigate the heat island effect and reduce risks from landslides and flooding (Krellenberg et al., 2013; Rodríguez Laredo, 2011). In the case of governance frameworks, there is clear evidence that incorporation of climate change considerations into wider city planning is still a challenge, as are more inter-sectoral and participative processes that have been linked to more effective policies (Barton, 2009; Barton, 2013; De Oliveira, 2009; Romero-Lankao et al., 2013a). Several metropolitan adaptation plans have been generated over the last five years, e.g. Bogotá, Buenos Aires, Esmeraldas, Quito, and Sao Paulo, although for the most part they have been restricted to the largest conglomerations and are often included as an addition to mitigation plans (Carmin et al., 2009; Luque et al., 2013; Romero-Lankao, 2007b; Romero-Lankao et al., 2013a; Romero-Lankao et al., 2012b). Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 27.3.6. Renewable Energy 27.3.6.1. Observed and Projected Impacts and Vulnerabilities Table 27-6 shows the relevance of RE in the Latin America energy matrix as compared to the world for 2009 according to the International Energy Agency statistics (IEA, 2012). Hydropower is the most representative source of RE and therefore analyzed separately from this section and all other RE sources (see case study in section 27.6.1.). Geothermal energy will be not discussed as it is assumed that there is no impact of climate change on the effectiveness of this energy type (Arvizu et al., 2011). [INSERT TABLE 27-6 HERE Table 27-6: Comparison of consumption of different energetics in Latin America and the world (in thousand tons of oil equivalent (ktoe) on a net calorific value basis).] Hydro, wind energy and biofuel production might be sensitive to climate change in Brazil (Lucena et al., 2009). With the vital role that RE plays in mitigating the effects of GCC, being by far the most important sources of non- hydro RE in SA and CA, this sensitivity demands the implementation of RE projects that will increase knowledge on the crops providing bioenergy. For historical reasons, CA and SA developed sugarcane as bioenergy feedstock. Brazil accounts for the most intensive RE production as bioethanol, which is used by the majority of the cars in the country (Goldemberg, 2008) whereas biodiesel comprises 5% of all diesel nationwide. With the continent s long latitudinal length, the expected impacts of climate change on plants will be complex due to a wide variety of climate conditions, so that different crops would have to be used in different regions. In Brazil, most of the biodiesel comes from soybeans, but there are promising new sources such as palm oil (Lucena et al., 2009). The development of palm oil as well as soybean are important factors that induce land use change, with a potential to influence stability of forests and biodiversity in certain key regions in SA, such as the Amazon (section 27.2.2.1). Biofuels can help CA and SA to decrease emissions from energy production and use. However, RE might imply potential problems such as those related to positive net emissions of greenhouse gases, threats to biodiversity, an increase in food prices and competition for water resources (section 27.2.3), some of which can be reverted or attenuated (Koh and Ghazoul, 2008). For example, the sugarcane agro-industry in Brazil combusts bagasse to produce electricity, providing power for the bioethanol industry and increasing sustainability. The excess heat energy is then used to generate bioelectricity, thus allowing the biorefinery to be self-sufficient in energy utilization (Amorim et al., 2011; Dias et al., 2012). In 2005/2006 the production of bioelectricity was estimated to be 9.2 kWh per ton of sugarcane (Macedo et al., 2008), approximately 2% of Brazil s total energy generation production. Most bioenergy feedstocks at present in production in CA and SA are grasses. In the case of sugarcane, the responses to the elevation of CO2 concentration up to 720 ppmv have been shown to be positive in terms of biomass production and principally regarding water use efficiency (De Souza et al., 2008). The production of energy from renewable sources such as hydro- and wind power is greatly dependent on climatic conditions and therefore may be impacted in the future by GCC. Lucena et al. (2010a) suggests an increasing energy vulnerability of the poorest regions of Brazil to GCC together with a possible negative influence on biofuels production and electricity generation, mainly biodiesel and hydropower respectively. (JAM: we use CC and GCC, I believe that we use should use CC only). Expansion of biofuel crops in Brazil might cause both direct and indirect land use changes (e.g., biofuel crops replacing rangelands, which previously replaced forests) with the direct land use changes, according to simulation performed by Lapola et al. (2010) of the effects for 2020. The same study shows that sugarcane ethanol and biodiesel derived from soybean each contribute with about one half of the indirect deforestation projected for 2020 (121.970 km2) (Lapola et al., 2010). Thus, indirect land use changes, especially those causing the rangeland frontier to move further into the Amazonian forests, might potentially offset carbon savings from biofuels production. Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The increase in global ethanol demand is leading to the development of new hydrolytic processes capable of converting cellulose into ethanol (Dos Santos et al., 2011). The expected increase in the hydrolysis technologies is very likely to balance the requirement of land for biomass crops. Thus, the development of these technologies has a strong potential to diminish social (e.g. negative health effects due the burning process, poor labor conditions) and environmental impacts (e.g. loss of biodiversity, water and land uses) whereas at it can improve the economic potential of sugarcane. One adaptation measure will be to increase the productivity of bioenergy crops due to planting in high productivity environments with highly developed technologies, in order to use less land. As one of the main centers of biotech agriculture application in the world (Gruskin, 2012), the region has a great potential to achieve this goal. As the effects previously reported on crops growing in SESA might prevail (see 27.3.4.1), i.e. that an increase in productivity may happen due to increasing precipitation, future uncertainty will have to be dealt with by preparing adapted varieties of soybean in order to maintain food and biodiesel production, mainly in Argentina as it is one of the main producers of biodiesel from soybean in the world (Chum et al., 2011). Other renewable energy sources such as wind power generation may also be vulnerable, raising the need for further research. According to Lucena et al. (2009; 2010b) the projections of changes in wind power in Brazil, may favor the use of this kind of energy in the future. 27.3.6.2. Adaptation Practices RE will become increasingly more important over time as this is closely related with the emissions of GHG (Fischedick et al., 2011). Thus, RE could have an important role as adaptation means to provide sustainable energy for development in the region (see also 27.6.1). However, the production of RE requires large available areas for agriculture, which is the case of Argentina, Bolivia, Brazil, Chile, Colombia, Peru and Venezuela, that together represent 90% of the total area of CA and SA. However, for small countries it might not be possible to use bioenergy. Instead, they could benefit in the future from other types of RE, such as geothermal, eolic, photovoltaic etc., depending on policies and investment in different technologies. This is important because economic development is thought to be strongly correlated with an increase in energy use (Smil, 2000), which is itself associated with an increase in emissions (Sathaye et al., 2011). Latin America is second to Africa in terms of technical potential for bioenergy production from rain-fed lignocellulosic feedstocks on unprotected grassland and woodlands (Chum et al., 2011). Among the most important adaptation measures regarding RE are: (1) management of land use change (LUC); and (2) development of policies for financing and management of science and technology for all types of RE in the region. If carefully managed, biofuel crops can be used as a means to regenerate biodiversity as proposed by Buckeridge et al. (2012) highlighting that the technology for tropical forest regeneration has become available and that forests could share land with biofuel crops (such as sugarcane) taking advantage of forests mitigating potential. A possible adaptation measure could be to expand the use of reforestation technology to other countries in CA and SA. One of the main adaptation issues is related to the food vs. fuel issues (Valentine et al., 2012). This is important because an increase in bioenergy feedstocks might threaten primary food production in a scenario expected to feed future populations with an increase of 70% in production (Gruskin, 2012; Valentine et al., 2012). This is particularly important in the region as it has one of the highest percentages of arable land available for food production in the world (Nellemann et al., 2009). As CA and SA develop new strategies to produce more RE there might be a pressure for more acreage to produce bioenergy. Because climate change will affect bioenergy and food crops at the same time, their effects, as well as the adaptation measures related to agriculture will be similar. The main risks identified by Arvizu et al. (2011) are: (1) business as usual; (2) un-reconciled growth, and (3) environment and food vs. fuel. Thus, the most important adaptation measures will be the ones related to the control of economic growth, environmental management and agriculture production. The choice for lignocellulosic feedstocks (e.g. sugarcane second generation technologies) will be an important mitigation/adaptation measure because these feedstocks do not compete with food (Arvizu et al., 2011). In the case of sugarcane, for instance, an increase of ca. 40% in the Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 production of bioethanol is expected as a result of the implantation of second generation technologies coupled with the first generation ones already existent in Brazil (De Souza et al., 2013; Dias et al., 2012). Biodiesel production has the lowest costs in Latin America (Chum et al., 2011) due to the high production of soybean in Brazil and Argentina. The use of biodiesel to complement oil-derived diesel is a productive choice for adaptation measures regarding this bioenergy source. Also, the cost of ethanol, mainly derived from sugarcane, is the lowest in CA, SA and Latin America (Chum et al., 2011) and as an adaptation measure, such costs, as well as the one of biodiesel, should be lowered even more by improving technologies related to agricultural and industrial production of both. Indeed, it has been reported that in LA the use of agricultural budgets by governments for investment in public goods induces faster growth, decreasing poverty and environmental degradation (López and Galinato, 2007). The pressure of soy expansion due to biodiesel demand can lead to land use change and consequently to economic teleconnections, as suggested by Nepstad et al. (2006). These teleconnections may link Amazon deforestation derived from soy expansion to the economic growth in some developing countries due to changes in the demand of soy. These effects may possibly mean a decrease in jobs related to small to big farms in agriculture in Argentina (Tomei and Upham, 2009) on the one hand, and deforestation in the Amazon due to the advance of soybean cropping in the region on the other (Nepstad and Stickler, 2008). 27.3.7. Human Health 27.3.7.1. Observed and Projected Impacts and Vulnerability Changes in weather extremes and climatic patterns are affecting human health (high confidence), by increasing morbidity, mortality, and disabilities, and through the emergence of diseases in previously non-endemic regions (high confidence) (Rodríguez-Morales, 2011; Winchester and Szalachman, 2009). Heat waves and cold spells have increased urban mortality rates (Bell et al., 2008; Hajat et al., 2010; Hardoy and Pandiella, 2009; McMichael et al., 2006; Muggeo and Hajat, 2009). Outbreaks of vector- and water-borne diseases were triggered in CA by hurricane Mitch in 1998 (Costello et al., 2009; Rodríguez-Morales et al., 2010), while the 2010-2012 Colombian floods caused hundreds of deaths and thousands of displaced people (Hoyos et al., 2013). The number of cases of malaria have increased in Colombia during the last five decades alongside air temperatures (Arevalo-Herrera et al., 2012; Poveda et al., 2011), but also in urban and rural Amazonian regions undergoing large environmental changes (Cabral et al., 2010; Da Silva-Nunes et al., 2012; Gil et al., 2007; Tada et al., 2007). Malaria transmission has reached 2,300 m in the Bolivian Andes, and vectors are found at higher altitudes from Venezuela to Bolivia (Benítez and Rodríguez-Morales, 2004; Lardeux et al., 2007; Pinault and Hunter, 2011). Although the incidence of malaria has decreased in Argentina, its vector density has increased in the northwest along with climate variables (Dantur Juri et al., 2011; Dantur Juri et al., 2010). El Nino drives malaria outbreaks in Colombia (Mantilla et al., 2009; Poveda et al., 2011) amidst other factors (Osorio et al., 2007; Restrepo-Pineda et al., 2008; Rodríguez- Morales et al., 2006). Linkages between ENSO and malaria are also reported in Ecuador and Peru (Anyamba et al., 2006; Kelly-Hope and Thomson, 2010), French Guiana (Hanf et al., 2011), Amazonia (Olson et al., 2009), and Venezuela (Moreno et al., 2007). Unlike malaria, dengue fever (DF) and its hemorrhagic variant (DHF) are mostly urban diseases whose vector is affected by climate conditions. Their incidence have risen in tropical America in the last 25 years, causing annual economic losses of US$ 2.1+[1 to 4] billion (Shepard et al., 2011; Tapia-Conyer et al., 2009; Torres and Castro, 2007). Environmental and climatic variability affect their incidence in CA (Fuller et al., 2009; Mena et al., 2011; Rodríguez-Morales et al., 2010), in Colombia (Arboleda et al., 2009), and in French Guiana alongside malaria (Carme et al., 2009; Gharbi et al., 2011). In Venezuela, DF increases during La Nina (Herrera-Martinez and Rodríguez-Morales, 2010; Rodríguez-Morales and Herrera-Martinez, 2009). Weather and climate variability are also associated with DF in southern SA (Costa et al., 2010; De Carvalho-Leandro et al., 2010; Degallier et al., 2010; Honório et al., 2009; Lowe et al., 2011), involving also demographic and geographic factors in Argentina (Carbajo et al., 2012). In Rio de Janeiro a 1°C increase in monthly minimum temperature led to a 45% increase of DF in the next month, and 10 mm increase in rainfall to a 6% increase (Gomes et al., 2012). Despite large vaccination Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 campaigns, the risk of Yellow Fever (YF) outbreaks has increased mostly in tropical America s densely populated poor urban settings (Gardner and Ryman, 2010), alongside climate conditions (Jentes et al., 2011). Schistosomiasis (SCH) is endemic in rural areas of Suriname, Venezuela, the Andean highlands, and rural and peripheral urbanized regions of Brazil (Barbosa et al., 2010; Igreja, 2011; Kelly-Hope and Thomson, 2010). It is highly likely that SCH will increase in a warmer climate (Lopes et al., 2010; Mangal et al., 2008; Mas-Coma et al., 2009). Vegetation indices are associated with human fascioliasis in the Andes (Fuentes, 2004). Hantaviruses (HV) have been recently reported throughout the region (Jonsson et al., 2010; MacNeil et al., 2011), and El Nino and climate change augment their prevalence (Dearing and Dizney, 2010). Variation in HV reservoirs in Patagonia is strongly dependent on climate and environmental conditions (Andreo et al., 2012; Carbajo et al., 2009). In Venezuela, Rotavirus (RV) is more frequent and more severe in cities with minimal seasonality (Kane et al., 2004). The peak of RV in Guatemala occurs in the dry season, causing 60% of total diarrhoea cases (Cortes et al., 2012). In spite of its rapid decline, climate-sensitive Chagas disease is still a major public health issue (Abad-Franch et al., 2009; Araújo et al., 2009; Gottdenker et al., 2011; Moncayo and Silveira, 2009; Tourre et al., 2008). Climate also affects the most prevalent mycosis (Barrozo et al., 2009), and ENSO is associated with outbreaks of bartonellosis in Peru (Payne and Fitchett, 2010). The high incidence of cutaneous leishmaniasis (CL) in Bolivia is exacerbated during La Nina (García et al., 2009; Gomez et al., 2006). CL is affected in Costa Rica by temperature, forest cover and ENSO (Chaves et al., 2008), and in Colombia by land cover, altitude, climatic variables, and El Nino (Cárdenas et al., 2006; 2007; 2008; Valderrama- Ardila et al., 2010), and decreases during La Nina in Venezuela (Cabaniel et al., 2005). CL in Suriname peaks during the March dry season (35%) (Van der Meide et al., 2008), and in French Guiana is intensified after the October-December dry season (Rotureau et al., 2007). Incidence of visceral leishmaniasis (VL) has increased in Brazil (highest in LA) in association with El Nino and deforestation (Cascio et al., 2011; Ready, 2008; Sortino- Rachou et al., 2011), as in Argentina, Paraguay, and Uruguay (Bern et al., 2008; Dupnik et al., 2011; Fernández et al., 2012; Salomón et al., 2011). VL transmission in Venezuela is associated with rainfall seasonality (Feliciangeli et al., 2006; Rodríguez-Morales et al., 2007). Besides, the incidence of skin cancer in Chile has increased in recent years, concomitantly with climate and geographic variables (Salinas et al., 2006). Onchocerciasis (river blindness) vector exhibits seasonal biting rates (Botto et al., 2005; Rodríguez-Pérez et al., 2011), and leptospirosis is prevalent in CA s warm-humid tropical regions (Valverde et al., 2008). Other climate- driven infectious diseases are ascariasis and gram-positive cocci in Venezuela (Benítez et al., 2004; Rodríguez- Morales et al., 2010), and Carrion s disease in Peru (Huarcaya et al., 2004). Sea water temperature affects the abundance of cholera s bacteria (Hofstra, 2011; Jutla et al., 2010; Koelle, 2009; Marcheggiani et al., 2010), which explains the outbreaks during El Nino in Peru, Ecuador, Colombia and Venezuela (Cerda Lorca et al., 2008; Gavilán and Martínez-Urtaza, 2011; Holmner et al., 2010; Martínez-Urtaza et al., 2008; Murugaiah, 2011; Salazar-Lindo et al., 2008). The worsening of air quality and higher temperatures in urban settings are increasing chronic respiratory and cardiovascular diseases, and morbidity from asthma and rhinitis (Grass and Cane, 2008; Gurjar et al., 2010; Jasinski et al., 2011; Martins and Andrade, 2008; Rodriguez et al., 2011), but also artherosclerosis, pregnancy-related outcomes, cancer, cognitive deficit, otitis, and diabetes (Olmo et al., 2011). Dehydration from heatwaves increases hospitalizations for chronic kidney diseases (Kjellstrom et al., 2010), affecting construction, sugarcane and cotton workers in CA (Crowe et al., 2009; 2010; Kjellstrom and Crowe, 2011; Peraza et al., 2012). Extreme weather/climate events affect mental health in Brazil (depression, psychological distress, anxiety, mania and bipolar disorder), in particular in drought-prone areas of NEB (Coelho et al., 2004; Volpe et al., 2010). Extreme weather, meager crop yields, and low GDP are also associated with increased violence (McMichael et al., 2006). Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Multiple factors increase the region s vulnerability to climate change: precarious health systems, malnutrition, inadequate water and sanitation services, poor waste collection and treatment systems, air, soil and water pollution, lack of social participation, and inadequate governance (Luber and Prudent, 2009; Rodríguez-Morales, 2011; Sverdlik, 2011). Human health vulnerabilities in the region depend on geography, age (Graham et al., 2011; Martiello and Giacchi, 2010; Perera, 2008; Astrom et al., 2011), gender (Oliveira et al., 2011), race, ethnicity, and socio-economic status (Diez Roux et al., 2007; Martiello and Giacchi, 2010). Neglected Tropical Diseases in LA cause 1.5-5.0 million disability-adjusted life years (DALYs) (Hotez et al., 2008). Vulnerability of mega-cities (see 27.3.5) is aggravated by access to clean water, rapid spread of diseases (Borsdorf and Coy, 2009), and migration from rural areas forced by disasters (Borsdorf and Coy, 2009; Campbell-Lendrum and Corvalán, 2007; Hardoy and Pandiella, 2009). Human health vulnerabilities have been assessed in Brazil through composite indicators involving downscaled climate scenarios, epidemiological variables and socio-economic projections (Barbieri and Confalonieri, 2011; Confalonieri et al., 2009; ; FIOCRUZ, 2011). The Andes and CA are among the regions of highest predicted losses [1% to 27%] in labor productivity from future climate scenarios (Kjellstrom et al., 2009). 27.3.7.2. Adaptation Strategies and Practices Adaptation efforts in the region ((Blashki et al., 2007; Costello et al., 2011) are hampered by lack of political commitment, gaps in scientific knowledge, and institutional weaknesses (Keim, 2008; Lesnikowski et al., 2011; Olmo et al., 2011) (see 27.4.3). Research priorities and current strategies must be reviewed (Halsnaes and Verhagen, 2007; Karanja et al., 2011; Romero and Boelaert, 2010), and preventive/responsive systems must be put in place (Bell, 2011) to foster adaptive capacity (Campbell-Lendrum and Bertollini, 2010; Huang et al., 2011). Colombia established a pilot adaptation program to cope with changes in malaria transmission and exposure (Poveda et al., 2011). The city of Sao Paulo has implemented local pollution control measures, with the co-benefit of reducing GHG emissions (De Oliveira, 2009; Nath and Behera, 2011). Human wellbeing indices must be explicitly stated as adaptation policies in LA (e.g., Millennium Development Goals) (Franco-Paredes et al., 2007; Halsnaes and Verhagen, 2007; Mitra and Rodriguez-Fernandez, 2010). South- south cooperation and multidisciplinary research are required to design relevant adaptation and mitigation strategies (Team and Manderson, 2011; Tirado et al., 2010). 27.4. Adaptation Opportunities, Constraints, and Limits 27.4.1. Adaptation Needs and Gaps During the last years, the study of adaptation to climate change has progressively switched from an impact-focused approach (mainly climate-driven) to include a vulnerability-focused vision (Boulanger et al., 2011). As a consequence, the development and implementation of systemic adaptation strategies, involving institutional, social, ecosystem, environmental, financial and capacity components (see Chapter 14), to cope with present climate extreme events is a key step toward climate change adaptation, especially in SA and CA countries. While different frameworks and definitions of vulnerability exist, a general tendency aims at studying vulnerability to climate change especially in SA and CA focusing on the following aspects: urban vulnerability (e.g. Hardoy and Pandiella, 2009; Heinrichs and Krellenberg, 2011), rural community (McSweeney and Coomes, 2011; Ravera et al., 2011), rural farmer vulnerability (Oft, 2010), and sectoral vulnerability (see 27.3). The approach used can be holistic or systemic (Carey et al., 2012b; Ison, 2010), where climate drivers are actually few with respect to all other drivers related to human and environment interactions including physical, economic, political and social context, as well as local characteristics such as occupations, resource uses, accessibility to water, etc. (Manuel-Navarrete et al., 2007; Young et al., 2010). In developing and emergent countries, there exists a general consensus that the adaptive capacity is low, strengthened by the fact that poverty is key determinant of vulnerability in Latin America (to climate related natural hazards, see Rubin and Rossing, 2012) and thus a limit to resilience (Pettengell, 2010) leading to a low human development trap (UNDP, 2007). However, Magnan (2009) suggests that this analysis is biased by a relative Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 immaturity of the science of adaptation to explain what are the processes and the determinants of adaptive capacity . Increasing research efforts on the study of adaptation is therefore of great importance to improve our understanding of the actual societal, economical, community and individual drivers defining the adaptive capacity. Especially, a major focus on traditions and their transmission (Young and Lipton, 2006) may actually indicate potential adaption potentials in remote and economically poor regions of SA and CA. Such a potential does not dismiss the fact that the nature of future challenges may actually not be compared to past climate variability (e.g. glacier retreat in the Andes). Coping with new situations may require new approaches such as a multilevel risk governance (Corfee-Morlot et al., 2011; Young and Lipton, 2006) associated with decentralization in decision-making and responsibility. While the multilevel risk governance and the local participatory approach are interesting frameworks for strengthening adaptation capacity, perception of local and national needs is diverging, challenging the implementation of adaptation strategies in CA/SA (Salzmann et al., 2009). At present, despite an important improvement during the last years, there still exists a certain lack of awareness of environmental changes and mainly their implications for livelihoods and businesses (Young et al., 2010). Moreover, considering the limited financial resources of some states in CA and SA, long-term planning and the related human and financial resource needs may be seen as conflicting with present social deficit in the welfare of the population. This situation weakens the importance of adaptation planning to climate change in the political agenda (Carey et al., 2012b), and requires therefore international involvement as one facilitating factor in natural hazard management and climate change adaptation, in accordance with the respect to sovereignty, the international conventions including the United Nations Framework Convention on Climate Change. In addition, as pointed out by McGray et al. (2007), development, adaptation and mitigation issues are not separate issues. Especially, development and adaptation strategies should be tackled together in developing countries such as SA and CA, focusing on strategies to reduce vulnerability. The poor level of adaptation of present-day climate in SA and CA countries is characterized by the fact that responses to disasters are mainly reactive rather than preventive. Some early warning systems are being implemented, but the capacity of responding to a warning is often limited, particularly among poor populations. Finally, actions combining public communication (and education), public decision-maker capacity-building and a synergetic development-adaptation funding will be key to sustain the adaptation process that CA and SA require to face future climate change challenges. 27.4.2. Practical Experiences of Autonomous and Planned Adaptation, including Lessons Learned Adaptation processes in many cases have been initiated a few years ago, and there is still a lack of literature to evaluate their efficiency in reducing vulnerability and building resilience of the society against climate change. However, experiences of effective adaptation and maladaptation are slowly being documented (see also 27.4.3), some lessons have already been learned from these first experiences (see section 27.3), and tools, such as the Index of Usefulness of Practices for Adaptation (IUPA) to evaluate adaptation practices have been developed for the region (Debels et al., 2009). Evidenced by these practical experiences, there is a wide range of options to foster adaptation and thus adaptive capacity in CA and SA. In CA and SA, many societal issues are strongly connected to development goals and are often considered priority in comparison to adaptation efforts to climate change. However, according to the 135 case studies analyzed by McGray et al. (2007), 21 of which were in CA and SA, the synergy between development and adaptation actions allows to ensuring a sustainable result of the development projects. Vulnerability and disaster risk reduction may not always lead to long-term adaptive capacity (Nelson and Finan, 2009; Tompkins et al., 2008), except when structural reforms based on good governance (Tompkins et al., 2008) and negotiations (Souza Filho and Brown, 2009) are implemented. While multi-level governance can help to create resilience and reduce vulnerability (Corfee-Morlot et al., 2011; Roncoli, 2006; Young and Lipton, 2006), capacity- building (Eakin and Lemos, 2006), good governance and enforcement (Lemos et al., 2010; Pittock, 2011) are key components. Autonomous adaptation experience are mainly realized at local levels (individual or communitarian) with examples found for instance for rural communities in Honduras (McSweeney and Coomes, 2011), indigenous communities in Bolivia (Valdivia et al., 2010) and coffee agroforestry systems in Brazil (De Souza et al., 2012). However, such Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 adaptation processes do not always respond specifically to climate forcing. For instance, the agricultural sector adapts rapidly to economic stressors, while, despite a clear perception of climate risks, it may last longer before responding to climate changes (Tucker et al., 2010). In certain regions or communities, such as Anchioreta in Brazil (Bonatti et al., 2012), adaptation is part of a permanent process and is actually tackled through a clear objective of vulnerability reduction, maintaining and diversifying a large set of natural varieties of corn allowing the farmers to diversify their planting. Another kind of autonomous adaptation is the southward displacement of agriculture activities (e.g. wine, coffee) though the purchase of lands, which will become favorable for such agriculture activities in a warmer climate. In Argentina, the increase of precipitation observed during the last 30 years contributed to a westward displacement of the annual crop frontier. However, local adaptation to climate and non- climate drivers may undermine long-term resilience of socio-ecological systems when local, short-term strategies designed to deal with specific threats or challenges do not integrate a more holistic and long-term vision of the system at threat (Adger et al., 2011). Thus, policy should identify the sources of and conditions for local resilience and strengthen their capacities to adapt and learn (Adger et al., 2011; Borsdorf and Coy, 2009; Eakin et al., 2011), as well as to integrate new adapted tools (Oft, 2010). This sets the question of convergence between the local- scale/short-term and broad scale/long-term visions in terms of perceptions of risks, needs to adapt and appropriate policies to be implemented (Eakin and Wehbe, 2009; Salzmann et al., 2009). Even if funding for adaptation is available, the overarching problem is the lack of capacity and/or willingness to address the risks, especially those threatening lower income groups (Satterthwaite, 2011a). Adaptation to climate change cannot eliminate the extreme weather risks, and thus efforts should focus on disaster preparedness and post-disaster response (Sverdlik, 2011). Migration is the last resort for rural communities facing water stress problems in CA and SA (Acosta-Michlik et al., 2008). In natural hazard management contributing to climate change adaptation, specific cases such as the one in Lake 513 in Peru (Carey et al., 2012b) clearly allowed to identify facilitating factors for a successful adaptation process (technical capacity, disaster events with visible hazards, institutional support, committed individuals, and international involvement) as well as impediments divergent risk perceptions, imposed government policies, institutional instability, knowledge disparities, and invisible hazards). In certain cases, forward-looking learning (anticipatory process), as a contrast to learning by shock (reactive process), has been found as a key element for adaptation and resilience (Tschakert and Dietrich, 2010) and should be promoted as a tool for capacity-building at all levels (stakeholders, local and national governments). Its combination with role-playing game and agent-based models (Rebaudo et al., 2011) can strengthen and accelerate the learning process. Planned adaptation policies promoted by governments have been strengthened by the participation in international networks, where experience and knowledge can be exchanged. As an example, the C40 Cities- Climate Leadership Group or ICLEI include Bogota (Colombia), Buenos Aires (Argentina), Caracas (Venezuela), Curitiba, Rio de Janeiro and Sao Paulo (Brazil), Lime (Peru) and Santiago de Chile (Chile). Most of these cities have come up with related action and strategy plans (e.g. Action Plan Buenos Aires 2030, Plan of Caracas 2020 or the Metropolitan Strategy to CCA of Lima) (C40 Cities, 2011). At a regional policy level, an example of intergovernmental initiatives in SA and CA is the Ibero-American Programme on Adaptation to Climate Change (PIACC), developed by the Ibero-American Network of Climate Change Offices (RIOCC) (Keller et al., 2011b). For CA specifically, the Central American Commission for Environment and Development (CCAD) brings together the environmental ministries of the Central American Integration System (Sistema de la Integración Centroamericana (SICA)) that released its climate change strategy in 2010 (CCAD-SICA, 2010; Keller et al., 2011a). These initiatives demonstrate that there has been a growing awareness of CA and SA governments on the need to integrate climate change and future climate risks in their policies. Up to date, in total 18 regional Non-Annex countries, including Argentina, Belize, Bolivia, Brazil, Chile, Colombia, Costa Rica, Ecuador, El Salvador, Guatemala, Honduras, Nicaragua, Guyana, Panama, Paraguay, Peru, Suriname, Uruguay and Venezuela, have already published their first and/or second National Communication to the UNFCCC (see UNFCCC, 2012) allowing to measure the country s emissions and to assess its present and future vulnerability. Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 27.4.3. Observed and Expected Barriers to Adaptation Adaptation is a dynamic process, which to be efficient requires a permanent evolution and even transformation of the vulnerable system. Such transformation process can be affected by several constraints, including constraints affecting the context of adaptation as well as the implementation of policies and measures (see 16.3.2). Major constraints related to the capacity and resources needed to support the implementation of adaptation policies and processes comprehend access to (Lemos et al., 2010) and exchange of knowledge (e.g. adaptive capacity can be enhanced by linking indigenous knowledge and scientific knowledge; Valdivia, 2010), the access to and quality of natural resources (López-Marrero, 2010), the access to financial resources, especially for poor households (Hickey and Weis, 2012; Rubin and Rossing, 2012; Satterthwaite, 2011b) as well as for institutions (Pereira et al., 2009), technological resources (López-Marrero, 2010), technical assistance (Eakin et al., 2011; Guariguata, 2009) , as well as the fostering of public-private technology transfer (La Rovere et al., 2009a; Ramirez-Villegas et al., 2012) and promotion of technical skills (Hickey and Weis, 2012), and social asset-based formation at the local level (Rubin and Rossing, 2012). In terms of framing adaptation, as constraint to affect the adaptation context, it is usually considered that a major barrier to adaptation is the perception of risks and many studies focused on such an issue (Bonatti et al., 2012). Also, new studies (Adger et al., 2009) identified social limits to possible adaptation to climate change in relation with issues of values and ethics, risk, knowledge and culture, even though such limits can evolve in time. Indeed, while being a necessary condition, perception may not be the main driver for initiating an adaptation process. As pointed out by Tucker et al. (2010) with a specific focus on CA, exogenous factors (economic, land tenure, cost, etc.) may actually strongly constrain the decision-making process involved in possible adaptation process. In that sense, efficient governance and management are key components in the use of climate and non-climate information in the decision-making and adaptation process. As a consequence, it is difficult to describe adaptation without defining at which level it is thought. Indeed, while a lot of efforts are invested in national and regional policy initiatives, most of the final adaptation efforts will be local. National and international (transborder) governance is key to build adaptive capacity (Engle and Lemos, 2010) and therefore to strengthen (or weaken) local adaptation through efficient policies and delivery of resources. At a smaller scale (Agrawal, 2008), local institutions can strongly contribute to vulnerability reduction and adaptation. However, at all levels, the efficiency in national and local adaptation activities strongly depend on the capacity-building and information transmission to decision-makers (Eakin and Lemos, 2006). 27.5 Interactions between Adaptation and Mitigation Synergies between adaptation and mitigation strategies on the local level can be reached due to self-organization of communities in cooperatives (see The SouthSouthNorth Capacity Building Module on Poverty Reduction (see SSN, 2006), which manages recycling or renewable energy production, leading to an increase in energy availability, thus production capacity and therefore new financial resources. Moreover, Venema and Cisse (2004) support also the development of decentralized renewable energy solutions for the growth of renewable energy in CA and SA (see also section 27.3.6) next to large infrastructure project (see their case studies for Argentina and Brazil). In spite of their smaller size (individual or communitarian), these solutions offer adaptation and mitigation benefits. On one hand, fossil-based energy consumption is reduced, while energy availability is increased. On the other hand, reduction of energy precariousness is key in any development strategy. Thus, it allows local community and individuals to growing socially and economically; and therefore to reducing its vulnerability avoiding the poverty trap (UNDP, 2007), and to initiating an adaptation process based on non-fossil fuel energy sources. Such initiatives also depend on local and organizational leaderships (UN-Habitat, 2011). Such integrated strategies of income generation as adaptation measures as well as production of renewable energy are also identified for vulnerable, small farmers diversifying their crops towards crops for vegetable oil and biodiesel production in Brazil. Barriers identified concern capacity building and logistical requirements making policy tools, credit mechanism, and organization into cooperatives, and fostered research necessary (La Rovere et al., 2009b). Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Other promising interactions of mitigation and adaptation are identified e.g. for the management of Brazilian tropical natural and planted forest (Guariguata, 2009). At national and regional scales, CA and SA countries will require the allocation of human and financial resources to adapt to climate change. While resources are limited, too large an economic dependence of these countries to fossil fuels will reduce their adaptive capacity. The reduction in energy consumption and the integration of renewable energies in their energetic matrix is therefore a key issue for all these countries in order to sustain their development and growth and therefore increase their adaptive capacity (see also section 27.3.6). Reforestation and avoided deforestation are important practices that contribute to both mitigation and adaptation efforts in the region as in other parts of the world. Maintaining forest cover can provide a suite of environmental services including local climate regulation, water regulation, reduced soil erosion, all of which can reduce the vulnerability of communities to variable climate (see section 27.3.2.2 on Ecosystem-based adaptation) (Vignola et al., 2009). 27.6. Case Studies 27.6.1. Hydropower Hydropower is the main source of renewable energy in CA and SA (see section 27.3.6). The region is only second to Asia in terms of hydropower energy generation in the world, displaying a 20% share of total annual generation and an average regional capacity factor of over 50% (Table 5-1 SRREN; IPCC, 2011). As a result, the region has by far the largest proportion of electricity generated through hydropower facilities (Table 27-6 in section 27.3.6.1). The hydropower proportion of total electricity production is over 40% in the region, and in some cases is near or close to 80%, as in the case of Brazil, Colombia and Costa Rica (IEA, 2012). Although there is debate, especially in tropical environments, about GHG emissions from hydropower reservoirs (Fearnside and Pueyo, 2012), this form of electricity generation is often seen as a major contributor to mitigating GHG emissions worldwide (see IPCC SRREN [5]; Kumar et al., 2011). But on the other hand, hydropower is a climate-related sector, thus making it prone to the potential effects of changing climate conditions (see section 27.3.1.1). In this regards the CA and SA region constitute a unique example to study these relations between climate change mitigation and adaptation in relation to hydropower generation. Diverse studies have analyzed the potential impacts of climate change on hydropower generation (see details in Table 27-4 in section 27.3.1.1). Maurer et al. (2009) studied future conditions for the Lempa River (El Salvador, Honduras and Guatemala) showing a potential reduction in hydropower capacity of 33% to 53% by 2070-2099 (Maurer et al., 2009). A similar loss is expected for the Sinu-Caribe basin in Colombia were, despite a general projection of increased precipitation, losses due to evaporation enhancement reduces inflows to hydroelectric systems, thus reducing electricity generation up to 35% (Ospina-Norena et al., 2009a). Further studies (Ospina- Norena et al., 2011a; 2011b) have estimated vulnerability indices for the hydropower sector in the same basin, and identified reservoir operation strategies to reduce this vulnerability. Overall reductions in hydropower generation capacity are also expected in Chile for the main hydropower generation river basins: Maule, Laja and Biobio (ECLAC, 2009a; McPhee et al., 2010; Stehr et al., 2010), and also in the Argentinean Limay River basin (Seoane and López, 2007). Ecuador, on the other hand, faces an increase in generation capacity associated with an increment in precipitation on its largest hydroelectric generation Paute River basin (Buytaert et al., 2010). Brazil, although being the country with the largest installed hydroelectric capacity in the region, still has unused generation capacity in sub-basins of the Amazon River (Soito and Freitas, 2011). However, future climate conditions plus environmental concerns pose an important challenge for the expansion of the system (Andrade et al., 2012; Finer and Jenkins, 2012; Freitas and Soito, 2009). According to Lucena et al. (2009), hydropower systems in southern Brazil (most significantly the Parana River system) could face a slight increase in energy production under an A2 scenario. However, the rest of the country s hydropower system, and especially those located in NEB, could face a reduction in power generation, thus reducing the reliability of the whole system (Lucena et al., 2009). An obvious implication of the mentioned impacts is the need to replace the energy lost through alternative (see 27.3.6.2) or traditional sources. Adaptation measures have been studied for Brazil (Lucena et al., 2010a), with Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 results implying an increase in natural gas and sugarcane bagasse electricity generation in the order of 300 TWh, increase in operation costs in the order of 7 billion USD annually and 50 billion USD in terms of investment costs by 2035. In Chile, the study by ECLAC (2009a) assumed that the loss in hydropower generation, in the order of 18 TWh for the 2011-2040 period (a little over 10% of actual total hydropower generation capacity) would be compensated by the least operating cost source available, coal-fired power plant, implying an increase of 2 MTCO2e of total GHG emissions (emissions for the electricity sector in Chile totaled 25 MTCO2e in 2009). Ospina-Norena (2011a; 2011b) studied some adaptation options, such as changes in water use efficiency or demand growth that could mitigate the expected impacts on hydropower systems in the Colombian Sinú-Caribe River basin. Changes in seasonality and total availability could also increase complexities in the management of multiple-use dedicated basins in Peru (Condom et al., 2012; Juen et al., 2007), Chile (ECLAC, 2009a), and Argentina (Seoane and López, 2007), that could affect the relationship between different water users within a basin. It is worth noting that those regions which are projected to face an increase in streamflow and associated generation capacity, such as Ecuador or Costa Rica, also share difficulties in managing deforestation, erosion and sedimentation which limits the useful life of reservoirs (see section 27.3.1.1). In these cases it is important to consider these effects in future infrastructure operation (Ferreira and Teegavarapu, 2012) and planning, and also to enhance the on-going process of recognizing the value of the relation between ecosystem services and hydropower system operations (Leguía et al., 2008) (see more on PES in section 27.3.2.2 and 27.6.2). 27.6.2. Payment for Ecosystem Services Payment for ecosystem services (PES) is commonly described as a set of transparent schemes for securing a well- defined ecosystem service (or a land use capable to secure that service) through conditional payments or compensations to voluntary providers (Engel et al., 2008; Tacconi, 2012). Van Noordwijk et al. (2012) provides a broader definition to PES by arguing that it encompasses three complementary approaches, (i) the one above, i.e., commodification of pre-defined ecosystem services so that prices can be negotiated between buyers and sellers; plus (ii) compensation for opportunities forgone voluntarily or by command and control decisions; and (iii) coinvestment in environmental stewardships. Therefore, the terms ´conservation agreements´, ´conservation incentives´ and ´community conservation´ are often used as synonyms or as something different or broader than PES (Cranford and Mourato, 2011; Milne and Niesten, 2009). For simplicity, we refer to PES in its broadest sense (sensu Van Noordwijk et al., 2012). Services subjected to such types of agreements often include regulation of freshwater flows, carbon storage, provision of habitat for biodiversity, and scenic beauty (De Koning et al., 2011; Montagnini and Finney, 2011). Since the ecosystems that provide the services are mostly privately owned, policies often aim at supporting landowners to maintain the provision of services over time (Kemkes et al., 2010). Irrespective of the debate of as to whether payments or compensations should be designed to focus on actions or results (Gibbons et al., 2011), experiences in Colombia, Costa Rica and Nicaragua show that PES can finance conservation, ecosystem restoration, and better land use practices (Montagnini and Finney, 2011; see also Table 27-5). However, based on examples from Ecuador and Guatemala, Southgate et al. (2010) argue that uniformity of payment for beneficiaries can be inefficient if recipients accept less compensation in return for conservation measures, or if recipients that promote greater environmental gains receive only the prevailing payment. Other setbacks to PES schemes might include cases where there is a perception of commoditization of nature and its intangible values (e.g. Bolivia, Cuba, Ecuador and Venezuela); other cases where mechanisms are inefficient to reduce poverty; slowness to build trust between buyers and sellers, as well as gender and land tenure issues that might arise (Asquith et al., 2008; Balvanera et al., 2012; Peterson et al., 2010; Van Noordwijk et al., 2012). Table 27-7 lists selected examples of PES schemes in Latin America, but a more complete and detailed list is given in Balvanera et al. (2012). [INSERT TABLE 27-7 HERE Table 27-7: Cases of government-funded PES schemes in CA and SA.] Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The PES concept (or ´fishing agreements´) also applies to coastal and marine areas, although only a few cases have been reported. Begossi (2011) argues that this is due to three factors: origin (the mechanism was originally designed for forests), monitoring (marine resources such as fish are more difficult to monitor than terrestrial resources) and definition of resource boundaries in offshore water. One example of a compensation mechanism in the region is the so-called defeso, in Brazil. It consists of a period (reproductive season) when fishing is forbidden by the government and fishermen receive a financial compensation. It applies to shrimp, lobster and both marine and freshwater fisheries (Begossi et al., 2011). 27.7. Data and Research Gaps The scarcity and difficult availability of high resolution high quality and continuous climate, oceanic and hydrological data, together with the availability of only very few complete regional studies, pose challenges for the region to address changes in climate variability and the identification of trends in extremes, in particular for CA. This situation hampers studies on frequency and variability of extremes, as well as impacts and vulnerability analyses of the present and future climates, and the development of vulnerability assessments and adaptation actions. Related to observed impacts in most sectors, there is an imbalance in information availability among countries. While more studies have been performed for Brazil, Southern SA and SESA region, much less are available for CA and for some regions of tropical SA. An additional problem is poor dissemination of results in peer-reviewed publications since most information is available only as grey literature. There is a need for studies focused on current impacts and vulnerabilities across sectors throughout CA and SA, with emphasis on extremes to improve risk management assessments. The complex interactions between climate and non-climate drivers difficult the assessment of impacts and projections, as is the case for water availability and streamflows owing to current and potential deforestation; overfishing and pollution regarding the impacts on fisheries, or impacts on hydroenergy production. The lack of interdisciplinary integrated studies limits our understanding of the complex interactions between natural and socio- economic systems. In addition, accelerating deforestation and land use changes, as well as changes in economic conditions, impose a continuous need for updated and available data sets that feed basic and applied studies. To address the global challenge of food security and food quality, both important issues in CA and SA, investment in scientific agricultural knowledge need to be reinforced, mainly with regard to the integration of agriculture with organic production; and the integration of food and bioenergy production. It is necessary to consider ethical aspects when the competition for food and bioenergy production is analyzed to identify which activity is most important at a given location and time and whether bioenergy production would affect food security for a particular population. Sea level rise and coastal erosion are also relevant issues; the lack of comparable measurements of sea level rise in CA and SA difficult the present and future integrated assessment of the impacts of SLR in the region. Of local and global importance will be improving our understanding of the physical oceanic processes, in particular of the Humboldt Current system flowing along the west coast of SA, being this one of the most fish productive system worldwide. More information and research about the impacts of climate variability and change on human health is needed. One problem is the difficulty in accessing health data that are not always archived and ready to be used in integrated studies. Another need refers to building the necessary critical mass of transdisciplinary scientists to tackle the climate change-human health problems in the region. The prevailing gaps in scientific knowledge hamper the implementation of adaptation strategies, thus demanding a review of of research priorities towards better disease control. With the aim of further studying the health impacts of climate change and identifying resilience, mitigation and adaptation strategies, South-South cooperation and multidisciplinary research are considered to be relevant priorities. In spite of the uncertainty that stems from global and regional climatic projections, the region needs to act in preparation of a possible increase in climate variability and in extremes. It is necessary to undertake research Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 activities leading to public policies to assist societies in coping with current climate variability, as for example, risk assessment and risk management. Other important aspect since the AR4 is the improvement of climate modeling and the generation of high-resolution climate scenarios, that in countries in CA and SA resulted in the first integrated regional studies on impacts and vulnerability assessments of climate change focusing on sectors such as agriculture, energy and human health. Research on adaptation and the scientific understanding of the various processes and determinants of adaptive capacity is also mandatory for the region, with particular emphasis on increasing adaptation capacity involving the traditional knowledge of ancestral cultures and how this knowledge is transmitted. Linking indigenous knowledge with scientific knowledge is important. Although some adaptation processes have been initiated in the recent years, there is no literature assessing their efficiency so far. The research agenda needs to address vulnerability and foster adaptation in the region; encompassing an inclusion of the regions researchers and focusing also on governance structures and action-oriented research that addresses resource distribution inequities. Regional and international partnerships, and research networks and programs have allowed linking those programs with local strategies for adaptation and mitigation, also providing opportunities to address research gaps and exchange among researchers. Examples are the European Union funded projects CLARIS LPB in SESA and AMAZALERT in Amazonia. Other important initiatives come from the IAI, WHO, GEF, IDB, ECLAC (CEPAL), La Red and BirdLife International, among others. The same holds for local international networks such as ICLEI or C40, of which CA and SA cities form part. The weADAPT initiative is a good example on how CA and SA practitioners, researchers and policy makers can have access to credible, high quality information and to share experiences and lessons learnt in other regions of the world. 27.8. Conclusions CA and SA harbor unique ecosystems and maximum biodiversity, with a variety of eco-climatic gradients rapidly changing from development initiatives. Agricultural and beef production as well as bioenergy crops are on the raise mostly by expanding agricultural frontiers. Poverty and inequality are decreasing, but at a low pace. Socioeconomic development shows a high level of heterogeneity and a very unequal income distribution, resulting in high vulnerability to climatic conditions. There is still a high and persistent level of poverty in most countries (45% for CA and 30% for SA for year 2010) in spite of the sustained economic growth observed in the last decade. The IPCC AR4 and SREX reports contain ample evidence of increase in extreme climate events in CA and SA. During 2000-2010, 630 weather and climate extreme events lead to 16,000 fatalities, and 46.6 million people affected, with estimated losses of US$ 208 million. During 2000-2009, 39 hurricanes occurred in the CA-Caribbean basin compared to 15 and 9 in the decade of 1980 and 1990, respectively. In SESA, more frequent and intense rainfall extremes have favored an increase in the occurrence of flash floods and landslides. In Amazonia extreme droughts were reported in 2005 and 2010 and record floods were observed in 2009 and 2012. In 2012-2013 an extreme drought affected NEB. While warming occurred in most of CA and SA, cooling was detected off the coast of Southern Peru and Chile. There is growing evidence that Andean glaciers (both tropical and extratropical) are retreating in response to warming trends. Increases in precipitation were registered in SESA, CA and the NAMS regions, while decreases were observed in Southern Chile, and a slight decrease in NEB after the middle 1970s. In CA it has been observed a gradual delay of the beginning of the rainfall season. SLR varied from 2 to 7 mm/yr between 1950 and 2008 in CA and SA, which is a reason for concern since a large proportion of the population of the region lives by the coast. Land use and land cover change are key drivers of regional environmental change in SA and CA. Natural ecosystems are affected by climate variability/change and land use change. Deforestation, land degradation, and biodiversity loss are mainly attributed to increased extensive agriculture for traditional export activities and bioenergy crops. Agricultural expansion has affected fragile ecosystems, causing severe environmental degradation Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and reducing the environmental services provided by these ecosystems. Deforestation has intensified the process of land degradation, increasing the vulnerability of communities exposed floods, landslides and droughts. Plant species are rapidly declining in CA and SA, with a high percentage of rapidly declining amphibian species. However, the region has still large extensions of natural vegetation cover, with the Amazon being the main example. Ecosystem- based Adaptation practices, such as the establishment of protected areas and their effective management, conservation agreements, community management of natural areas, and payment for ecosystem services are increasingly more common across the region. Figure 27-7 summarizes of some of the main observed trends in global environmental change drivers across different representative regions of CA and SA. Changes in climate and non-climate drivers have to be compounded with other socioeconomic related trends; such as the rapid urbanization process experienced the region Some observed impacts on human and natural systems can be directly or indirectly attributed to human influences, and can be summarized as (Figure 27-8): Changes in river flow variability in the Amazon River during the last two decades, and robust positive trends in streamflow in sub-basins of the La Plata River basin, and increased dryness for most of the river basins in west coast of South America during the last 50 years. Reduction in tropical glaciers and icefields in tropical and extra tropical Andes over the second half of the 20th century that can be attributed to an increase in temperature Coastal erosion, bleaching of coral reefs in the coast of CA, and reduction in fisheries stock. Increase in agricultural yield in SESA, and shifting in agricultural zoning: significant expansion of agricultural areas, mainly in climatically marginal regions. Increase in frequency and extension of dengue fever, yellow fever and malaria. However, for some impacts the number of concluding studies is still insufficient, leading to low levels of confidence for attribution to human influences. [INSERT FIGURE 27-7 HERE Figure 27-7: Summary of observed changes in climate and other environmental factors in representative regions of CA and SA. The boundaries of the regions in the map are conceptual (neither geographic nor political precision). Information and references to changes provided are presented in different sections of the chapter.] [INSERT FIGURE 27-8 HERE Figure 27-8: Observed impacts of climate variations and attribution of causes in CA and SA.] By the end of the century, the CMIP5 derived projections for RCP8.5 projected: CA: mean annual warming of 2.5C (range: 1.5°C to 5.0 °C), mean rainfall reduction of 10% (range: -25% to +10%), and reduction in summertime precipitation; SA: mean warming of 4C (range: 2.0°C to 5.0 °C), with rainfall reduction up to 15% in tropical SA east of the Andes, and an increase of about 15-20% in SESA, and in other regions of the continent. Increases in warm days and nights are very likely to occur in most of SA. SESA: Increases in heavy precipitation. Northeastern South America: increases in dry spell. However, there is some degree of uncertainty on climate change projections for regions, particularly for rainfall in CA and tropical SA. Current vulnerability in terms of water supply in the semi-arid zones and the tropical Andes is expected to increase even further due to climate change. This would be exacerbated by the expected glacier retreat, precipitation reduction and increase evapotranspiration demands as expected in the semi-arid regions of CA and SA. These scenarios would affect water supply for large cities, small communities, hydropower generation and food production. There is a need for re-assessing current practices to reduce the mismatch between water supply and demand in order to reduce future vulnerability, and to implement constitutional and legal reforms towards more efficient and effective water resources management. SLR due to climate change and human activities on coastal and marine ecosystems pose threats to fish stocks, corals, mangroves, recreation and tourism, and diseases control in CA and SA. Coral reefs, mangroves, fisheries, and other benthic marine invertebrates that provide key ecosystem services, such as nutrient cycling, water quality Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 regulation, and herbivory, are also threatened by climate change. It is possible that the Mesoamerican coral reef will collapse by mid-century (between 2050 and 2070), causing major economic and environmental losses. In the southwestern Atlantic coast, eastern Brazilian reefs might suffer a massive coral cover decline in the next 50 years. In the Rio de La Plata area extreme-flooding events may become more frequent since return periods are decreasing, and urban coastal areas in the eastern coast will be particularly affected. Beach erosion is expected to increase in southern Brazil and in scattered areas at the Pacific coast. Urban populations in CA and SA face diverse social, political, economic and environmental risks in daily life, and climate change will add a new dimension to these risks. Since urban development remains fragile in many cases, with weak planning responses, climate change can compound existing challenges, e.g. water supply in cities from glacier, snowmelt and paramos related runoff in the Andes (Lima, La Paz/El Alto, Santiago de Chile, Bogota), flooding in several cities like Sao Paulo and Buenos, and health related challenges in many cities of the region. Climate change will affect individual species and biotic interactions. Vertebrate fauna will suffer major species losses especially in high altitude areas; elevational specialists might be particularly vulnerable because of their small geographic ranges and high energetic requirements; freshwater fisheries can suffer alterations in physiology and life histories. In addition, modifications in phenology, structure of ecological networks, predator-preys interactions and non-trophic interactions among organisms will affect biotic interactions. Shifts in biotic interactions are expected to have negative consequences on biodiversity and ecosystem services in High Andean ecosystems. Although in the region biodiversity conservation is largely confined to protected areas it is expected that many species and vegetational types will lose representativeness inside such protected areas. Changes in food production and food security are expected to have a great spatial variability, with a wide range of uncertainty mainly related to climate and crop models. In SESA average productivity could be sustained or increased until the mid-century, although interannual and decadal climate variability are likely to impose important damages. In other regions like NEB, CA, and some Andean countries agricultural productivity could decrease in the short-term, threatening the food security of the poorest population. The expansion of pastures and croplands is expected to continue in the coming years, particularly from an increasing global demand for food and biofuels. The great challenge for CA and SA will be to increase the food and bioenergy production and at the same time sustain the environmental quality in a scenario of climate change. Renewable energy provides a great potential for adaptation and mitigation. Hydropower is currently the main source of RE in CA and SA, followed by biofuels. SESA is one of the main sources of production of the feedstocks for biofuels production, mainly with sugarcane and soybean, and future climate conditions may lead to an increase in productivity and production. Advances in second generation biofuels will be important as a measure of adaptation, as they have the potential to increase biofuels productivity. In spite of the large amount of arable land available, the expansion of biofuels might have some direct and indirect land use change effects, producing teleconnections that could lead to deforestation of native tropical forests and loss of employment in some countries. This might also affect food security. Changes in weather and climatic patterns are negatively affecting human health in CA and SA, either by increasing morbidity, mortality, and disabilities, and through the emergence of diseases in previously non-endemic regions. Multiple factors increase the region s vulnerability to climate change: precarious health systems, malnutrition, inadequate water and sanitation services, population growth, poor waste collection and treatment systems, air, soil and water pollution, food in poor regions, lack of social participation, and inadequate governance. Vulnerabilities vary with geography, age, gender, race, ethnicity, and socio-economic status, and are rising in large cities. Climate change and variability may exacerbate current and future risks to health. Climate change will bring modifications to environmental conditions in space and time, and the frequency and intensity of weather and climate processes. In many CA and SA countries, a first step toward adaptation to climate change is to reduce the vulnerability to present climate, taking into account future potential impacts, particularly of weather and climate extremes. Long-term planning and the related human and financial resource needs may be seen as conflicting with present social deficit in the welfare of the CA and SA population. Such conditions weaken the importance of adaptation planning to climate change on the political agenda. Currently, there are few experiences on Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 synergies between development, adaptation and mitigation planning, which can help local communities and governments to allocate available resources in the design of strategies to reduce vulnerability and to develop adaptation measures. Facing a new climate system and, in particular, the exacerbation of extreme events, will call for new ways to manage human and natural systems for achieving sustainable development. [INSERT TABLE 27-8 HERE Table 27-8: Key risks from climate change and the potential for risk reduction through mitigation and adaptation.] Frequently Asked Questions FAQ 27.1: What is the impact of glacier retreat on natural and human systems in the tropical Andes? [to be inserted at end of Section 27.3.1.2] The retreat of glaciers in the tropical Andes mountains, with some fluctuations, started after the Little Ice Age (16th to 19th centuries), but the rate of retreat (area reduction between 20-50%) has accelerated since the late 1970s. the changes in runoff from glacial retreat into the basins fed by such runoff vary depending on the size and phase of glacier retreat. In an early phase, runoff tends to increase due to accelerated melting, but after a peak, as the glacierized water reservoir gradually empties, runoff tends to decrease. This reduction in runoff is more evident during dry months when glacier melt is the major contribution to runoff (high confidence). A reduction in runoff could endanger high Andean wetlands (bofedales) and intensify conflicts between different water users among the highly vulnerable populations in high elevation Andean tropical basins. Glacier retreat has also been associated with disasters such as glacial lake outburst floods that are a continuous threat in the region. Glacier retreat could also impact activities in high mountainous ecosystems such as alpine tourism, mountaineering and adventure tourism (high confidence). FAQ 27.2: Can payment for ecosystem services (PES) be used as an effective way to help local communities adapt to climate change? [to be inserted in Section 27.3.3.2] Ecosystems provide a wide range of basic services, like providing breathable air, drinkable water, and moderating flood risk (very high confidence). Assigning values to these services and designing conservation agreements based on these (broadly known as PES), can be an effective way to help local communities adapt to climate change. It can simultaneously help protect natural areas, and improve livelihoods and human well-being (medium confidence). However, during design and planning, a number of factors need to be taken into consideration at the local level in order to avoid potentially negative results. Problems can arise if a) the plan sets poor definitions about whether the program should focus just on actions to be taken or the end result of those actions, b) many perceive the initiative as commoditization of nature and its intangible values, c) the action is inefficient to reduce poverty, d) difficulties emerge in building trust between various stakeholders involved in agreements, and e) there are eventual gender or land tenure issues. FAQ 27.3: Are there emerging and re emerging human diseases as a consequence of climate variability and change in the region? [to be inserted in Section 27.3.7.2] Human health impacts have been exacerbated by variations and changes in climate extremes. Climate-related diseases have appeared in previously non-endemic regions (e.g. malaria in the Andes, dengue in CA and Southern SA) (high confidence). Climate variability and air pollution have also contributed to increase the incidence of respiratory and cardiovascular, vector- and water-borne and chronic kidney diseases, Hantaviruses and rotaviruses, pregnancy-related outcomes, and psychological trauma (very high confidence). Health vulnerabilities vary with geography, age, gender, ethnicity, and socio-economic status, and are rising in large cities. Without adaptation measures (e.g. extending basic public health services), climate change will exacerbate future health risks, owing to population growth rates and existing vulnerabilities in health, water, sanitation and waste collection systems, nutrition, pollution, and food production in poor regions (medium confidence). Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 References Abad-Franch, F., F.A. Monteiro, N. Jaramillo O., R. Gurgel-Gonçalves, F.B.S. Dias, and L. Diotaiuti, 2009: Ecology, evolution, and the long-term surveillance of vector-borne Chagas disease: A multi-scale appraisal of the tribe Rhodniini (Triatominae). Acta Tropica, 110(2-3), 159-177. Abell, R., M.L. Thieme, C. Revenga, M. Bryer, M. Kottelat, N. Bogutskaya, B. Coad, N. Mandrak, S. Contreras- Balderas, W. Bussing, M.L.J. Stiassny, P. Skelton, G.R. Allen, P. Unmack, A. Naseka, R. Ng, N. Sindorf, J. Robertson, E. Armijo, J.V. Higgins, T.J. Heibel, E. Wikramanayake, D. Olson, H.L. López, R.E. Reis, J.G. Lundberg, M.H.S. Pérez, and P. Petry, 2008: Freshwater ecoregions of the world: a new map of biogeographic units for freshwater biodiversity conservation. Bioscience, 58, 403-414. Abers, R.N., 2007: Organizing for Governance: Building Collaboration in Brazilian River Basins. World Development, 35(8), 1450-1463. Abson, D.J. and M. Termansen, 2011: Valuing Ecosystem Services in Terms of Ecological Risks and Returns. Conservation Biology, 25(2), 250-258. Acosta-Michlik, L., U. Kelkar, and U. Sharma, 2008: A critical overview: Local evidence on vulnerabilities and adaptations to global environmental change in developing countries. Global Environmental Change, 18(4), 539- 542. Adger, W.N., S. Dessai, M. Goulden, M. Hulme, I. Lorenzoni, D.R. Nelson, L.O. Naess, J. Wolf, and A. Wreford, 2009: Are there social limits to adaptation to climate change? Climatic Change, 93(3-4), 335-354. Adger, W.N., K. Brown, D.R. Nelson, F. Berkes, H. Eakin, C. Folke, K. Galvin, L. Gunderson, M. Goulden, K. O'Brien, J. Ruitenbeek, and E.L. Tompkins, 2011: Resilience implications of policy responses to climate change. Wiley Interdisciplinary Reviews-Climate Change, 2(5), 757-766. Aerts, J.C.J.H., H. Renssen, P.J. Ward, H. de Moel, E. Odada, L.M. Bouwer, and H. Goosse, 2006: Sensitivity of global river discharges under Holocene and future climate conditions . Geophysical Research Letters, 33(19), L19401. Agrawal, A., 2008: The Role of Local Institutions in Adaptation to Climate Change. In: IFRI Working Paper. Paper prepared for the Social Dimensions of Climate Change, Social Development Department, The World Bank, Washington DC. School of Natural Resources and Environment University of Michigan, pp. 47pp. Aguayo, M., A. Pauchard, G. Azócar, and O. Parra, 2009: Cambio del uso del suelo en el centro sur de Chile a fines del siglo XX. Entendiendo la dinámica espacial y temporal del paisaje. Revista Chilena De Historia Natural, 82, 361-374. Aguilar, M.Y., T.R. Pacheco, J.M. Tobar, and J.C. Quinonez, 2009: Vulnerability and adaptation to climate change of rural inhabitants in the central coastal plain of El Salvador. Climate Research, 40(2-3), 187-198. Aide, T.M., M.L. Clark, H.R. Grau, D. López-Carr, M.A. Levy, D. Redo, M. Bonilla-Moheno, G. Riner, M.J. Andrade-Núnez, and M. Muniz, 2013: Deforestation and Reforestation of Latin America and the Caribbean (2001-2010). Biotropica, 45(2), 262-271. Allison, E.H., A.L. Perry, M.-. Badjeck, W. Neil Adger, K. Brown, D. Conway, A.S. Halls, G.M. Pilling, J.D. Reynolds, N.L. Andrew, and N.K. Dulvy, 2009: Vulnerability of national economies to the impacts of climate change on fisheries. Fish and Fisheries, 10(2), 173-196. Alongi, D.M., 2008: Mangrove forests: Resilience, protection from tsunamis, and responses to global climate change. Estuarine Coastal and Shelf Science, 76(1), 1-13. Altieri, M. and P. Koohafkan, 2008: Enduring Farms: Climate Change, Smallholders and Traditional Farming Communities. In: TWN Environment and Development Series 6. Third World Network (TWN), Penang, Malaysia, pp. 63pp. Alves, M.d.C., L.G. de Carvalho, E.A. Pozza, L. Sanches, and J.C.d.S. Maia, 2011: Ecological zoning of soybean rust, coffee rust and banana black sigatoka based on Brazilian climate changes. Earth System Science 2010: Global Change, Climate and People, 6, 35-49. Amorim, H.V., M.L. Lopes, J.V.d. Castro Oliveira, M.S. Buckeridge, and G.H. Goldman, 2011: Scientific challenges of bioethanol production in Brazil. Applied Microbiology and Biotechnology, 91(5), 1267-1275. Amsler, M.L. and E.C. Drago, 2009: A review of the suspended sediment budget at the confluence of the Paraná and Paraguay Rivers. Hydrological Processes, 23(22), 3230-3235. Anciaes, M. and A.T. Peterson, 2006: Climate Change Effects on Neotropical Manakin Diversity Based on Ecological Niche Modeling. The Condor, 108(4), 778-791. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Anderson, B.T., J. Wang, G. Salvucci, S. Gopal, and S. Islam, 2010: Observed Trends in Summertime Precipitation over the Southwestern United States. Journal of Climate, 23(7), 1937-1944. Andrade, E.M., J.P. Cosenza, L.P. Rosa, and G. Lacerda, 2012: The vulnerability of hydroelectric generation in the Northeast of Brazil: The environmental and business risks for CHESF. Renewable & Sustainable Energy Reviews, 16(8), 5760-5769. Andrade, M.I. and O.E. Scarpati, 2007: Recent changes in flood risk in the Gran La Plata, Buenos Aires province, Argentina: causes and management strategy. GeoJournal, 70(4), 245-250. Andreo, V., C. Provensal, S. Levis, N. Pini, D. Enria, and J. Polop, 2012: Summer-autumn distribution and abundance of the hantavirus host, Oligoryzomys longicaudatus, in northwestern Chubut, Argentina. Journal of Mammalogy, 93(6), 1559-1568. Anthelme, F., B. Buendia, C. Mazoyer, and O. Dangles, 2012: Unexpected mechanisms sustain the stress gradient hypothesis in a tropical alpine environment. Journal of Vegetation Science, 23(1), 62-72. Anyamba, A., J.-. Chretien, J. Small, C.J. Tucker, and K.J. Linthicum, 2006: Developing global climate anomalies suggest potential disease risks for 2006 - 2007. International Journal of Health Geographics, 5. Araújo, C.A.C., P.J. Waniek, and A.M. Jansen, 2009: An overview of chagas disease and the role of triatomines on its distribution in Brazil. Vector-Borne and Zoonotic Diseases, 9(3), 227-234. Arboleda, S., N. Jaramillo-O., and A.T. Peterson, 2009: Mapping environmental dimensions of dengue fever transmission risk in the Aburrá Valley, Colombia. International Journal of Environmental Research and Public Health, 6(12), 3040-3055. Arevalo-Herrera, M., M.L. Quinones, C. Guerra, N. Céspedes, S. Giron, M. Ahumada, J.G. Pineros, N. Padilla, Z. Terrientes, A. Rosas, J.C. Padilla, A.A. Escalante, J.C. Beier, and S. Herrera, 2012: Malaria in selected non- Amazonian countries of Latin America. Acta Tropica, 121(3), 303-314. Arias, P.A., R. Fu, C.D. Hoyos, W. Li, and L. Zhou, 2011: Changes in cloudiness over the Amazon rainforests during the last two decades: diagnostic and potential causes. Climate Dynamics, 37(5-6), 1151-1164. Arias, P.A., R. Fu, and K.C. Mo, 2012: Decadal Variation of Rainfall Seasonality in the North American Monsoon Region and Its Potential Causes. Journal of Climate, 25(12), 4258-4274. Armenteras, D., E. Cabrera, N. Rodríguez, and J. Retana, 2013: National and regional determinants of tropical deforestation in Colombia. Regional Environmental Change, 1-13. Arnell, N.W. and S.N. Gosling, 2013: The impacts of climate change on river flow regimes at the global scale. Journal of Hydrology, 486, 351-364. Arvizu, D., T. Bruckner, H. Chum, O. Edenhofer, S. Estefen, A. Faaij, M. Fischedick, G. Hansen, G. Hiriart, O. Hohmeyer, K.G.T. Hollands, J. Huckerby, S. Kadner, A. Killingtveit, A. Kumar, A. Lewis, O. Lucon, P. Matschoss, L. Maurice, M. Mirza, C. Mitchell, W. Moomaw, J. Moreira, L.J. Nilsson, J. Nyboer, R. Pichs- Madruga, J. Sathaye, J.L. Sawin, R. Schaeffer, T.A. Schei, S. Schlömer, K. Seyboth, R. Sims, G. Sinden, Y. Sokona, C.v. Stechow, J. Steckel, A. Verbruggen, R. Wiser, F. Yamba, and T. Zwickel, 2011: Technical Summary. In: IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 158. Asquith, N.M., M.T. Vargas, and S. Wunder, 2008: Selling two environmental services: In-kind payments for bird habitat and watershed protection in Los Negros, Bolivia. Ecological Economics, 65(4), 675-684. Astrom, D.O., B. Forsberg, and J. Rocklov, 2011: Heat wave impact on morbidity and mortality in the elderly population: A review of recent studies. Maturitas, 69(2), 99-105. Ayoo, C., 2008: Economic instruments and the conservation of biodiversity. Management of Environmental Quality, 19(5), 550-564. Baethgen, W.E., 2010: Climate Risk Management for Adaptation to Climate Variability and Change. Crop Science, 50(Supplement 1), 70-76. Baker, A.C., P.W. Glynn, and B. Riegl, 2008: Climate change and coral reef bleaching: An ecological assessment of long-term impacts, recovery trends and future outlook. Estuarine, Coastal and Shelf Science, 80(4), 435-471. Baldi, G. and J.M. Paruelo, 2008: Land-Use and Land Cover Dynamics in South American Temperate Grasslands. Ecology and Society, 13(2), 6. Balvanera, P., M. Uriarte, L. Almeida-Lenero, A. Altesor, F. DeClerck, T. Gardner, J. Hall, A. Lara, P. Laterra, M. Pena-Claros, D.M. Silva Matos, A.L. Vogl, L.P. Romero-Duque, L.F. Arreola, Á.P. Caro-Borrero, F. Gallego, M. Jain, C. Little, R. de Oliveira Xavier, J.M. Paruelo, J.E. Peinado, L. Poorter, N. Ascarrunz, F. Correa, M.B. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cunha-Santino, A.P. Hernández-Sánchez, and M. Vallejos, 2012: Ecosystem services research in Latin America: The state of the art. Ecosystem Services, 2(0), 56-70. Baraer, M., B. Mark, J. McKenzie, T. Condom, J. Bury, K. Huh, C. Portocarrero, J. Gomez, and S. Rathay, 2012: Glacier recession and water resources in Peru s Cordillera Blanca. Journal of Glaciology, 58(207), 134-150. Barbieri, A. and U.E.C. Confalonieri, 2011: Climate Change, Migration and Health: exploring potential scenarios of population vulnerability in Brazil. In: Etienne Piguet; Antoine Pecoud (Org.). Migration and Climate Change. Cambridge University Press and UNESCO Publishing, Cambridge, UK, pp. 49-73. Barbosa, C.S., K.C. Araújo, M.A.A. Sevilla, F. Melo, E.C.S. Gomes, and R. Souza-Santos, 2010: Current epidemiological status of schistosomiasis in the state of Pernambuco, Brazil. Memorias do Instituto Oswaldo Cruz, 105(4), 549-554. Barcaza, G., M. Aniya, T. Matsumoto, and T. Aoki, 2009: Satellite-Derived Equilibrium Lines in Northern Patagonia Icefield, Chile, and Their Implications to Glacier Variations. Arctic Antarctic and Alpine Research, 41(2), 174-182. Bárcena, A., 2010: Structural Constraints on Development in Latin America and the Caribbean: A Post-Crisis Reflection. Cepal Review, 100, 7-27. Barros, V., 2008: Adaptation to Climate Trends: Lessons From the Argentine Experience. In: Climate Change and Adaptation. [Leary, N., J. Adejuwon, V. Barros, I. Burton, J. Kulkarm, and R. Lasco (eds.)]. Earthscan, London, UK and New York, USA, pp. 398pp. Barros, V., A. Menéndez, C. Natenzon, R. Kokot, J. Codignotto, M. Re, P. Bronstein, I. Camilloni, S. Luduena, and D. Rios, 2008: Storm Surges, Rising Seas and Flood Risks in Metropolitan Buenos Aires. In: Climate Change and Vulnerability. [Leary, N., C. Conde, J. Kulkarni, A. Nyong, and J. Pulhin (eds.)]. Earthscan, London, UK, pp. 117-132. Barros, V.R., 2010: El Cambio Climático en Argentina (Capitulo 3). In: Agro y Ambiente: una agenda compartida para el desarrollo sustentable. Foro de la Cadena Agroindustrial Argentina, Buenos Aires, Argentina, pp. 35pp. Barrozo, L.V., R.P. Mendes, S.A. Marques, G. Benard, M.E. Siqueira Silva, and E. Bagagli, 2009: Climate and acute/subacute paracoccidioidomycosis in a hyper-endemic area in Brazil. International Journal of Epidemiology, 38(6), 1642-1649. Barrucand, M.G., W.M. Vargas, and M.M. Rusticucci, 2007: Dry conditions over Argentina and the related monthly circulation patterns. Meteorology and Atmospheric Physics, 98(1-2), 99-114. Barton, J.R., 2009: Adaptación al cambio climático en la planificación de ciudades-regiones. Revista De Geografía Norte Grande, 43, 5-30. Barton, J.R., 2013: Climate Change Adaptive Capacity in Santiago de Chile: Creating a Governance Regime for Sustainability Planning. International Journal of Urban and Regional Research, published online. Bathurst, J.C., J. Amezaga, F. Cisneros, M. Gavino Novillo, A. Iroumé, M.A. Lenzi, J. Mintegui Aguirre, M. Miranda, and A. Urciuolo, 2010: Forests and floods in Latin America: science, management, policy and the EPIC FORCE project. Water International, 35(2), 114-131. Bathurst, J.C., S.J. Birkinshaw, F. Cisneros, J. Fallas, A. Iroumé, R. Iturraspe, M.G. Novillo, A. Urciuolo, A. Alvarado, C. Coello, A. Huber, M. Miranda, M. Ramirez, and R. Sarandón, 2011: Forest impact on floods due to extreme rainfall and snowmelt in four Latin American environments 2: Model analysis. Journal of Hydrology, 400(3-4), 292-304. Battisti, D.S. and R.L. Naylor, 2009: Historical Warnings of Future Food Insecurity with Unprecedented Seasonal Heat. Science, 323(5911), 240-244. Begossi, A., P.H. May, P.F. Lopes, L.E.C. Oliveira, V. da Vinha, and R.A.M. Silvano, 2011: Compensation for environmental services from artisanal fisheries in SE Brazil: Policy and technical strategies. Ecological Economics, 71, 25-32. Bell, A.R., N.L. Engle, and M.C. Lemos, 2011: How does diversity matter? The case of Brazilian river basin councils. Ecology and Society, 16(1), 42. Bell, E., 2011: Readying health services for climate change: A policy framework for regional development. American Journal of Public Health, 101(5), 804-813. Bell, M.L., M.S. O'Neill, N. Ranjit, V.H. Borja-Aburto, L.A. Cifuentes, and N.C. Gouveia, 2008: Vulnerability to heat-related mortality in Latin America: A case-crossover study in Sao Paulo, Brazil, Santiago, Chile and Mexico City, Mexico. International Journal of Epidemiology, 37(4), 796-804. Benayas, J.M.R., A.C. Newton, A. Diaz, and J.M. Bullock, 2009: Enhancement of Biodiversity and Ecosystem Services by Ecological Restoration: A Meta-Analysis. Science, 325(5944), 1121-1124. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Benegas, L., F. Jimenez, B. Locatdlli, J. Faustino, and M. Campos, 2009: A Methodological Proposal for the Evaluation of Farmer's Adaptation to Climate Variability, Mainly Due to Drought in Watersheds in Central America. Mitigation and Adaptation Strategies for Global Change, 14(2), 169. Benhin, J.K.A., 2006: Agriculture and deforestation in the tropics: A critical theoretical and empirical review. Ambio, 35(1), 9-16. Benítez, J., Rodríguez, A., Sojo, M., Lobo, H., Villegas, C., Oviedo, L., et al. (2004). Descripción de un brote epidémico de malaria de altura en un área originalmente sin malaria del estado Trujillo, Venezuela. Boletín de Malariología y Salud Ambiental, 44(2), 93-100. Benítez, J.A. and A.J. Rodríguez-Morales, 2004: Malaria de Altura en Venezuela ?Consecuencia de las variaciones climáticas? (in revision). CIMEL, 9(1), 27-30. Bern, C., J.H. Maguire, and J. Alvar, 2008: Complexities of assessing the disease burden attributable to leishmaniasis. PLoS Neglected Tropical Diseases, 2(10). Berthrong, S.T., E.G. Jobbágy, and R.B. Jackson, 2009: A global meta-analysis of soil exchangeable cations, pH, carbon, and nitrogen with afforestation. Ecological Applications, 19(8), 2228-2241. Bettolli, M.L., W.M. Vargas, and O.C. Penalba, 2009: Soya bean yield variability in the Argentine Pampas in relation to synoptic weather types: monitoring implications. Meteorological Applications, 16(4), 501-511. Betts, R.A., P.M. Cox, M. Collins, P.P. Harris, C. Huntingford, and C.D. Jones, 2004: The role of ecosystem- atmosphere interactions in simulated Amazonian precipitation decrease and forest dieback under global climate warming. Theoretical and Applied Climatology, 78(1-3), 157-175. Betts, R.A., Y. Malhi, and J.T. Roberts, 2008: The future of the Amazon: new perspectives from climate, ecosystem and social sciences. Philosophical Transactions of the Royal Society B-Biological Sciences, 363(1498), 1729- 1735. Blashki, G., T. McMichael, and D.J. Karoly, 2007: Climate change and primary health care. Australian Family Physician, 36(12), 986-989. Blázquez, J. and M.N. Nunez, 2013: Analysis of uncertainties in future climate projections for South America: comparison of WCRP-CMIP3 and WCRP-CMIP5 models. Climate Dynamics, 41(3-4), 1039-1056. Bodin, X., F. Rojas, and A. Brenning, 2010: Status and evolution of the cryosphere in the Andes of Santiago (Chile, 33.5 degrees S.). Geomorphology, 118(3-4), 453-464. Bombardi, R.J. and L.M.V. Carvalho, 2009: IPCC global coupled model simulations of the South America monsoon system. Climate Dynamics, 33(7-8), 893-916. Bonatti, M., E. Gentile, A.C.F.d. Vasconcelos, L.H.I. Ribeiro Homem, L.R. D'Agostini, and S.L. Schlindwein, 2012: Vulnerability to Climate Change and Different Perceptions of Social Actors: Thinking about motivation problems (Manuscript Draft). Climatic Change,. Borsdorf, A. and M. Coy, 2009: Megacities and global change: Case studies from Latin America. Erde, 140(4), 341- 353. Botto, C., E. Escalona, S. Vivas-Martinez, V. Behm, L. Delgado, and P. Coronel, 2005: Geographical patterns of onchocerciasis in southern Venezuela: Relationships between environment and infection prevalence. Parassitologia, 47(1), 145-150. Boulanger, J.-., S. Schlindwein, and E. Gentile, 2011: CLARIS LPB WP1: Metamorphosis of the CLARIS LPB European project: from a mechanistic to a systemic approach. In: CLIVAR Exchanges No. 57, Vol. 16, No.3. World Climate Research Programme (WCRP), pp. 7-10. Bown, F., A. Rivera, and C. Acuna, 2008: Recent glacier variations at the Aconcagua basin, central Chilean Andes. Annals of Glaciology, 48. Bown, F. and A. Rivera, 2007: Climate changes and recent glacier behaviour in the Chilean Lake District. Global and Planetary Change, 59(1-4), 79-86. Bradley, R.S., M. Vuille, H.F. Diaz, and W. Vergara, 2006: Threats to water supplies in the tropical Andes. Science, 312(5781), 1755-1756. Bradley, R.S., F.T. Keimig, H.F. Diaz, and D.R. Hardy, 2009: Recent changes in freezing level heights in the Tropics with implications for the deglacierization of high mountain regions. Geophysical Research Letters, 36(17). Bradshaw, C.J.A., N.S. Sodhi, and B.W. Brook, 2009: Tropical turmoil: a biodiversity tragedy in progress. Frontiers in Ecology and the Environment, 7(2), 79-87. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Broad, K., A. Pfaff, R. Taddei, A. Sankarasubramanian, U. Lall, and de Souza Filho,Franciso de Assis, 2007: Climate, stream flow prediction and water management in northeast Brazil: societal trends and forecast value. Climatic Change, 84(2), 217-239. Brook, B.W., N.S. Sodhi, and C.J.A. Bradshaw, 2008: Synergies among extinction drivers under global change. Trends in Ecology & Evolution, 23(8), 453-460. Brooker, R.W., F.T. Maestre, R.M. Callaway, C.L. Lortie, L.A. Cavieres, G. Kunstler, P. Liancourt, K. Tielboerger, J.M.J. Travis, F. Anthelme, C. Armas, L. Coll, E. Corcket, S. Delzon, E. Forey, Z. Kikvidze, J. Olofsson, F. Pugnaire, C.L. Quiroz, P. Saccone, K. Schiffers, M. Seifan, B. Touzard, and R. Michalet, 2008: Facilitation in plant communities: the past, the present, and the future. Journal of Ecology, 96(1), 18-34. Bucher, E.H. and E. Curto, 2012: Influence of long-term climatic changes on breeding of the Chilean flamingo in Mar Chiquita, Córdoba, Argentina. Hydrobiologia, 697(1), 127-137. Buckeridge, M.S., A.P. De Souza, R.A. Arundale, K.J. Anderson-Teixeira, and E.d. Lucia, 2012: Ethanol from sugarcane in Brazil: a midway' strategy for increasing ethanol production while maximizing environmental benefits. Global Change Biology Bioenergy, 4(2), 119-126. Bulte, E.H., R. Damania, and R. Lopez, 2007: On the gains of committing to inefficiency: Corruption, deforestation and low land productivity in Latin America. Journal of Environmental Economics and Management, 54(3), 277-295. Burte, J.D.P., A. Coudrain, and S. Marlet, 2011: Use of water from small alluvial aquifers for irrigation in semi-arid regions. Revista Ciencia Agronômica, 42, 635-643. Bury, J.T., B.G. Mark, J.M. McKenzie, A. French, M. Baraer, K.I. Huh, M.A.Z. Luyo, and R.J.G. Lopez, 2011: Glacier recession and human vulnerability in the Yanamarey watershed of the Cordillera Blanca, Peru. Climatic Change, 105(1-2), 179-206. Bury, J., B.G. Mark, M. Carey, K.R. Young, J.M. McKenzie, M. Baraer, A. French, and M.H. Polk, 2013: New Geographies of Water and Climate Change in Peru: Coupled Natural and Social Transformations in the Santa River Watershed. Annals of the Association of American Geographers, 103(2), 363-374. Butler, R. and W.F. Laurance, 2009: Is oil palm the next emerging threat to the Amazon? Tropical Conservation Science, 2(1), 1-10. Butt, N., P.A. de Oliveira, and M.H. Costa, 2011: Evidence that deforestation affects the onset of the rainy season in Rondonia, Brazil. Journal of Geophysical Research-Atmospheres, 116, D11120. Buytaert, W., M. Vuille, A. Dewulf, R. Urrutia, A. Karmalkar, and R. Célleri, 2010: Uncertainties in climate change projections and regional downscaling in the tropical Andes: implications for water resources management. Hydrology and Earth System Sciences Discussion, 14(7), 1821-1848. Buytaert, W. and B. De Bievre, 2012: Water for cities: The impact of climate change and demographic growth in the tropical Andes. Water Resources Research, 48(8), 13pp. Buytaert, W., F. Cuesta-Camacho, and C. Tobón, 2011: Potential impacts of climate change on the environmental services of humid tropical alpine regions. Global Ecology and Biogeography, 20(1), 19-33. C40 Cities, 2011: C40 Cities. Available at: http://www.c40cities.org/c40cities (Accessed on 12 February 2013). C40 Cities, Climate Leadership Group. Cabaniel, G., L. Rada, J.J. Blanco, A.J. Rodríguez-Morales, and J.P. Escalera A., 2005: Impacto de los eventos de El Nino Southern oscillation (ENSO) sobre la leishmaniosis cutánea en Sucre, Venezuela, a través del uso de información satelital, 1994 - 2003. Revista Peruana De Medicina Experimental y Salud Publica, 22(1), 32-37. Cabral, A.C., N.F. Fe, M.C. Suarez-Mutis, M.N. Boia, and F.A. Carvalho-Costa, 2010: Increasing incidence of malaria in the Negro River basin, Brazilian Amazon. Transactions of the Royal Society of Tropical Medicine and Hygiene, 104(8), 556-562. Cabré, M., S. Solman, and M. Nunez, 2010: Creating regional climate change scenarios over southern South America for the 2020 s and 2050 s using the pattern scaling technique: validity and limitations Springer Netherlands, , pp. 449-469. Cáceres, B., 2010: Actualización del inventario de tres casquetes glaciares del Ecuador. Diss. Master of Science, Univ. Nice, France, 84 pp. Callaway, R.M., 2007: Positive Interactions and Interdependence in Plant Communities. Springer, Dordrecht, The Netherlands, pp. 415. Calmon, M., P.H.S. Brancalion, A. Paese, J. Aronson, P. Castro, S.C. da Silva, and R.R. Rodrigues, 2011: Emerging Threats and Opportunities for Large-Scale Ecological Restoration in the Atlantic Forest of Brazil. Restoration Ecology, 19(2), 154-158. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Camargo, M.B.P., 2010: The impact of climatic variability and climate change on arabic coffee crop in Brazil. Bragantia, 69(1), 239-247. Campbell, J.D., M.A. Taylor, T.S. Stephenson, R.A. Watson, and F.S. Whyte, 2011: Future climate of the Caribbean from a regional climate model. International Journal of Climatology, 31(12), 1866-1878. Campbell-Lendrum, D. and C. Corvalán, 2007: Climate change and developing-country cities: Implications for environmental health and equity. Journal of Urban Health, 84(SUPPL. 1), i109-i117. Campbell-Lendrum, D. and R. Bertollini, 2010: Science, media and public perception: Implications for climate and health policies. Bulletin of the World Health Organization, 88(4), 242. Campos, J.N.B. and T.M.d. Carvalho Studart, 2008: Drought and water policies in Northeast Brazil: backgrounds and rationale. Water Policy, 10(5), 425. Carbajo, A.E., C. Vera, and P.L.M. Gonzalez, 2009: Hantavirus reservoir Oligoryzomys longicaudatus spatial distribution sensitivity to climate change scenarios in Argentine Patagonia. International Journal of Health Geographics, 8(44), 1-10. Carbajo, A.E., M.V. Cardo, and D. Vezzani, 2012: Is temperature the main cause of dengue rise in non-endemic countries? The case of Argentina. International Journal of Health Geographics, 11, 26. Cárdenas, R., C.M. Sandoval, A.J. Rodríguez-Morales, and C. Franco-Paredes, 2006: Impact of climate variability in the occurrence of leishmaniasis in Northeastern Colombia. American Journal of Tropical Medicine and Hygiene, 75(2), 273-277. Cárdenas, R., C.M. Sandoval, A.J. Rodriguez-Morales, and C. Franco-Paredes, 2007: Climate variability and leishmaniasis in Colombia. American Journal of Tropical Medicine and Hygiene, 77(5), 286-286. Cárdenas, R., C.M. Sandoval, A.J. Rodriguez-Morales, and P. Vivas, 2008: Zoonoses and climate variability: The example of leishmaniasis in southern departments of Colombia [Sparagano O.A.E., Maillard J.-C., and Figueroa J.V.(eds.)]. , pp. 326-330. Carey, M., 2005: Living and dying with glaciers: people's historical vulnerability to avalanches and outburst floods in Peru. Global and Planetary Change, 47(2-4), 122-134. Carey, M., A. French, and E. O'Brien, 2012a: Unintended effects of technology on climate change adaptation: an historical analysis of water conflicts below Andean Glaciers. Journal of Historical Geography, 38(2), 181-191. Carey, M., C. Huggel, J. Bury, C. Portocarrero, and W. Haeberli, 2012b: An integrated socio-environmental framework for glacier hazard management and climate change adaptation: lessons from Lake 513, Cordillera Blanca, Peru. Climatic Change, 112(3-4), 733-767. Carilli, J.E., R.D. Norris, B.A. Black, S.M. Walsh, and M. McField, 2009: Local Stressors Reduce Coral Resilience to Bleaching. PLoS ONE, 4(7), e6324. Carme, B., S. Matheus, G. Donutil, O. Raulin, M. Nacher, and J. Morvan, 2009: Concurrent dengue and malaria in cayenne hospital, French Guiana. Emerging Infectious Diseases, 15(4), 668-671. Carmin, J.A., D. Roberts, and I. Anguelovski, 2009: Planning Climate Resilient Cities: Early Lessons from Early Adapters. In: Paper presented at World Bank 5th Urban Research Symposium, Cities and Climate Change, Marseille, France, 28-30 June 2009, pp. 27pp. Carmona, A. and G. Poveda, 2011: Identificación de modos principales de variabilidad hidroclimática en Colombia mediante la transformada de Hilbert-Huang. In: IX Congreso Colombiano de Meteorología 25 March 2011, Auditorio Hemeroteca Nacional Bogotá, pp. 14. Carpenter, K.E., M. Abrar, G. Aeby, R.B. Aronson, S. Banks, A. Bruckner, A. Chiriboga, J. Cortés, J.C. Delbeek, L. DeVantier, G.J. Edgar, A.J. Edwards, D. Fenner, H.M. Guzmán, B.W. Hoeksema, G. Hodgson, O. Johan, W.Y. Licuanan, S.R. Livingstone, E.R. Lovell, J.A. Moore, D.O. Obura, D. Ochavillo, B.A. Polidoro, W.F. Precht, M.C. Quibilan, C. Reboton, Z.T. Richards, A.D. Rogers, J. Sanciangco, A. Sheppard, C. Sheppard, J. Smith, S. Stuart, E. Turak, J.E.N. Veron, C. Wallace, E. Weil, and E. Wood, 2008: One-Third of Reef-Building Corals Face Elevated Extinction Risk from Climate Change and Local Impacts. Science, 321(5888), 560-563. Carr, D.L., A. Carla Lopez, and R.E. Bilsborrow, 2009: The population, agriculture, and environment nexus in Latin America: country-level evidence from the latter half of the twentieth century. Population and Environment, 30(6), 222-246. Carrasco, J.F., G. Casassa, and J. Quintana, 2005: Changes of the 0°C isotherm and the equilibrium line altitude in central Chile during the last quarter of the 20th century / Changements de l'isotherme 0°C et de la ligne d'équilibre des neiges dans le Chili central durant le dernier quart du 20eme siecle. Hydrological Sciences Journal, 50(6). Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Carvalho, L.M.V., C. Jones, A.E. Silva, B. Liebmann, and P.L. Silva Dias, 2011: The South American Monsoon System and the 1970s climate transition. International Journal of Climatology, 31(8), 1248-1256. Casassa, G., W. Haeberli, G. Jones, G. Kaser, P. Ribstein, A. Rivera, and C. Schneider, 2007: Current status of Andean glaciers. Global and Planetary Change, 59(1-4), 1-9. Casassa, G., P. López, B. Pouyaud, and F. Escobar, 2009: Detection of changes in glacial run-off in alpine basins: examples from North America, the Alps, central Asia and the Andes. Hydrological Processes, 23(1), 31-41. Cascio, A., M. Bosilkovski, A.J. Rodriguez-Morales, and G. Pappas, 2011: The socio-ecology of zoonotic infections. Clinical Microbiology and Infection, 17(3), 336-342. Casco, S.L., C.E. Natenzon, N.I. Basterra, and J.J. Neiff, 2011: Inundaciones en el Bajo Paraná ?Se puede articular la gestión social a partir del comportamiento hidrológico previo? Interciencia, 36(6), 423-430. Castellanos, E.J., C. Tucker, H. Eakin, H. Morales, J.F. Barrera, and R. Díaz, 2013: Assessing the adaptation strategies of farmers facing multiple stressors: Lessons from the Coffee and Global Changes project in Mesoamerica. Environmental Science & Policy, 26, 19-28. Cavazos, T., C. Turrent, and D.P. Lettenmaier, 2008: Extreme precipitation trends associated with tropical cyclones in the core of the North American monsoon. Geophysical Research Letters, 35(21). CCAD-SICA, 2010: Estrategia Regional de Cambio Climático. Comisión Centroamericana de Ambiente y Desarrollo (CCAD)/ Sistema de la Integración Centroamericana (SICA), Antiguo Cuscatlán, El Salvador, Centroamérica, pp. 93pp. Ceballos, J.L., C. Euscátegui, J. Ramírez, M. Canon, C. Huggel, W. Haeberli, and H. Machguth, 2006: Fast shrinkage of tropical glaciers in Colombia. Annals of Glaciology, 43, 194-201. CEPALSTAT, 2012: Database and Statistical Publications. Available at: http://www.cepal.org/estadisticas/default.asp?idioma=IN ECLAC. Cerda Lorca, J., G. Valdivia C., M.T. Valenzuela B., and J. Venegas L., 2008: Climate change and infectious diseases. A novel epidemiological scenario. Revista Chilena De Infectologia, 25(6), 447-452. Chaves, L.F., J.M. Cohen, M. Pascual, and M.L. Wilson, 2008: Social Exclusion Modifies Climate and Deforestation Impacts on a Vector-Borne Disease. Plos Neglected Tropical Diseases, 2(2), e176. Chazdon, R.L., 2008: Beyond deforestation: Restoring forests and ecosystem services on degraded lands. Science, 320(5882), 1458-1460. Chazdon, R.L., C.A. Harvey, O. Komar, D.M. Griffith, B.G. Ferguson, M. Martínez-Ramos, H. Morales, R. Nigh, L. Soto-Pinto, M. van Breugel, and S.M. Philpott, 2009: Beyond Reserves: A Research Agenda for Conserving Biodiversity in Human-modified Tropical Landscapes. Biotropica, 41(2), 142-153. Chen, J.L., C.R. Wilson, B.D. Tapley, D.D. Blankenship, and E.R. Ivins, 2007: Patagonia icefield melting observed by gravity recovery and climate experiment (GRACE). Geophysical Research Letters, 34(22), L22501. Chevallier, P., B. Pouyaud, W. Suarez, and T. Condom, 2011: Climate change threats to environment in the tropical Andes: glaciers and water resources. Regional Environmental Change, 11(S1), 179-187. Christie, D.A., J.A. Boninsegna, M.K. Cleaveland, A. Lara, C. Le Quesne, M.S. Morales, M. Mudelsee, D.W. Stahle, and R. Villalba, 2011: Aridity changes in the Temperate-Mediterranean transition of the Andes since ad 1346 reconstructed from tree-rings. Climate Dynamics, 36(7-8), 1505-1521. Chum, H., A. Faaij, J. Moreira, G. Berndes, P. Dhamija, H. Dong, B. Gabrielle, A.G. Eng, W. Lucht, M. Mapako, O.M. Cerutti, T. McIntyre, T. Minowa, and K. Pingoud, 2011: Bioenergy. In: IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Chuvieco, E., S. Opazo, W. Sione, H. Del Valle, J. Anaya, C. Di Bella, I. Cruz, L. Manzo, G. Lopez, N. Mari, F. Gonzalez-Alonso, F. Morelli, A. Setzer, I. Csiszar, J. Ander Kanpandegi, A. Bastarrika, and R. Libonati, 2008: Global burned-land estimation in Latin America using modis composite data RID D-2396-2010. Ecological Applications, 18(1), 64-79. Coe, M.T., E.M. Latrubesse, M.E. Ferreira, and M.L. Amsler, 2011: The effects of deforestation and climate variability on the streamflow of the Araguaia River, Brazil. Biogeochemistry, 105(1-3), 119-131. Coe, M.T., M.H. Costa, and B.S. Soares-Filho, 2009: The influence of historical and potential future deforestation on the stream flow of the Amazon River Land surface processes and atmospheric feedbacks. Journal of Hydrology, 369(1 2), 165-174. Coelho, A.E.L., J.G. Adair, and J.S.P. Mocellin, 2004: Psycological responses to drought in Northeastern Brazil. Interamerican Journal of Psychology, 38(1), 95-103. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Collini, E.A., E.H. Berbery, V.R. Barros, and M.E. Pyle, 2008: How Does Soil Moisture Influence the Early Stages of the South American Monsoon? Journal of Climate, 21(2), 195-213. Condom, T., M. Escobar, D. Purkey, J.C. Pouget, W. Suarez, C. Ramos, J. Apaestegui, A. Tacsi, and J. Gomez, 2012: Simulating the implications of glaciers' retreat for water management: a case study in the Rio Santa basin, Peru. Water International, 37(4), 442-459. Confalonieri, U.E.C., D.P. Marinho, and R.E. Rodriguez, 2009: Public health vulnerability to climate change in Brazil. Climate Research, 40(2-3), 175-186. Confalonieri, U.E.C. and et al., 2011: Social, Environmental and Health Vulnerability to Climate Change in the Brazilian Northeastern Region . Climate Change, (submitted). Conway, D. and G. Mahé, 2009: River flow modelling in two large river basins with non-stationary behaviour: the Paraná and the Niger. Hydrological Processes, 23(22), 3186-3192. Cooper, E., L. Burke, and N. Bood, 2008: Belize's Coastal Capital: The Economic Contribution of Belize's Coral Reefs and Mangroves. Available at: http://www.wri.org/publications. In: WRI Working Paper. World Resources Institute (WRI), Washington DC, USA, pp. 53. Corfee-Morlot, J., I. Cochran, S. Hallegatte, and P. Teasdale, 2011: Multilevel risk governance and urban adaptation policy. Climatic Change, 104(1), 169-197. Cortés, G., X. Vargas, and J. McPhee, 2011: Climatic sensitivity of streamflow timing in the extratropical western Andes Cordillera. Journal of Hydrology, 405(1-2), 93-109. Cortes, J., W. Arvelo, B. Lopez, L. Reyes, T. Kerin, R. Gautam, M. Patel, U. Parashar, and K.A. Lindblade, 2012: Rotavirus disease burden among children <5years of age - Santa Rosa, Guatemala, 2007-2009. Tropical Medicine and International Health, 17(2), 254-259. Costa Fereira, L.d., R. D'Almeida Martins, F. Barbi, L.d. Costa Ferreira, L.F.d. Mello, A. Matenhauer Urbinatti, F.O.d. Souza, and T.H.N.d. Andrade, 2011: Governing Climate Change in Brazilian Coastal Cities: Risks and Strategies. Journal of US-China Public Administration, 8(1), 51-65. Costa, E.A.P.A., E.M.M. Santos, J.C. Correia, and C.M.R. de Albuquerque, 2010: Impact of small variations in temperature and humidity on the reproductive activity and survival of Aedes aegypti (Diptera, Culicidae). Revista Brasileira De Entomologia, 54(3), 488-493. Costa, L.C., F. Justino, L.J.C. Oliveira, G.C. Sediyama, W.P.M. Ferreira, and C.F. Lemos, 2009: Potential forcing of CO 2 , technology and climate changes in maize ( Zea mays ) and bean ( Phaseolus vulgaris ) yield in southeast Brazil. Environmental Research Letters, 4(1), 014013. Costa, M.H. and G.F. Pires, 2010: Effects of Amazon and Central Brazil deforestation scenarios on the duration of the dry season in the arc of deforestation. International Journal of Climatology, 30(13), 1970-1979. Costa, M.H., S.N.M. Yanagi, P.J.O.P. Souza, A. Ribeiro, and E.J.P. Rocha, 2007: Climate change in Amazonia caused by soybean cropland expansion, as compared to caused by pastureland expansion RID A-5695-2009. Geophysical Research Letters, 34(7), L07706. Costello, A., M. Abbas, A. Allen, S. Ball, S. Bell, R. Bellamy, S. Friel, N. Groce, A. Johnson, M. Kett, M. Lee, C. Levy, M. Maslin, D. McCoy, B. McGuire, H. Montgomery, D. Napier, C. Pagel, J. Patel, J.A.P. de Oliveira, N. Redclift, H. Rees, D. Rogger, J. Scott, J. Stephenson, J. Twigg, J. Wolff, and C. Patterson, 2009: Managing the health effects of climate change. Lancet and University College London Institute for Global Health Commission. The Lancet, 373(9676), 1693-1733. Costello, A., M. Maslin, H. Montgomery, A.M. Johnson, and P. Ekins, 2011: Global health and climate change: Moving from denial and catastrophic fatalism to positive action. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1942), 1866-1882. Cox, P.M., R.A. Betts, C.D. Jones, S.A. Spall, and I.J. Totterdell, 2000: Acceleration of global warming due to carbon-cycle feedbacks in a coupled climate model . Nature, 408(6809), 184-187. Cox, P.M., R.A. Betts, M. Collins, P.P. Harris, C. Huntingford, and C.D. Jones, 2004: Amazonian forest dieback under climate-carbon cycle projections for the 21st century . Theoretical and Applied Climatology, 78(1-3), 137-156. Cranford, M. and S. Mourato, 2011: Community conservation and a two-stage approach to payments for ecosystem services. Ecological Economics, 71, 89-98. CRED, 2011: EM-DAT. The International Disaster Database. Accessible at: http://www.emdat.be/. Collaborating Centre for Research on the Epidemiology of Disasters (CRED). Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cronkleton, P., M.R. Guariguata, and M.A. Albornoz, 2012: Multiple use forestry planning: Timber and Brazil nut management in the community forests of Northern Bolivia. Forest Ecology and Management, 268, 49-56. Crowe, J., B. van Wendel de Joode, and C. Wesseling, 2009: A pilot field evaluation on heat stress in sugarcane workers in Costa Rica: What to do next? Global Health Action, 2. Crowe, J., J. Manuel Moya-Bonilla, B. Roman-Solano, and A. Robles-Ramirez, 2010: Heat exposure in sugarcane workers in Costa Rica during the non-harvest season. Global Health Action, 3, 5619. Da Silva-Nunes, M., M. Moreno, J.E. Conn, D. Gamboa, S. Abeles, J.M. Vinetz, and M.U. Ferreira, 2012: Amazonian malaria: Asymptomatic human reservoirs, diagnostic challenges, environmentally driven changes in mosquito vector populations, and the mandate for sustainable control strategies. Acta Tropica, 121(3), 281-291. Dai, A., T. Qian, K.E. Trenberth, and J.D. Milliman, 2009: Changes in Continental Freshwater Discharge from 1948 to 2004. Journal of Climate, 22(10), 2773-2792. Dai, A., 2011: Drought under global warming: a review. Wiley Interdisciplinary Reviews-Climate Change, 2(1), 45- 65. DaMatta, F.M., A. Grandis, B.C. Arenque, and M.S. Buckeridge, 2010: Impacts of climate changes on crop physiology and food quality. Food Research International, 43(7), 1814-1823. Dantur Juri, M.J., M. Stein, and M.A. Mureb Sallum, 2011: Occurrence of Anopheles (Anopheles) neomaculipalpus Curry in north-western Argentina. Journal of Vector Borne Diseases, 48(1), 64-66. Dantur Juri, M.J., G.L. Claps, M. Santana, M. Zaidenberg, and W.R. Almirón, 2010: Abundance patterns of Anopheles pseudopunctipennis and Anopheles argyritarsis in northwestern Argentina. Acta Tropica, 115(3), 234-241. De Carvalho-Leandro, D., A.L.M. Ribeiro, J.S.V. Rodrigues, C.M.R. de Albuquerque, A.M. Acel, F.A. Leal-Santos, D.P. Leite Jr., and R.D. Miyazaki, 2010: Temporal distribution of Aedes aegypti Linnaeus (Diptera, Culicidae), in a Hospital in Cuiaba, State of Mato Grosso, Brazil. Revista Brasileira De Entomologia, 54(4), 701-706. De Koning, F., M. Aguinaga, M. Bravo, M. Chiu, M. Lascano, T. Lozada, and L. Suarez, 2011: Bridging the gap between forest conservation and poverty alleviation: the Ecuadorian Socio Bosque program. Environmental Science & Policy, 14(5), 531-542. De Mello, E.L., F.A. Oliveira, F.F. Pruski, and J.C. Figueiredo, 2008: Effect of the Climate Change on the Water Availability in the Paracatu River Basin. Engenharia Agricola, 28(4), 635-644. De Oliveira, J.A.P., 2009: The implementation of climate change related policies at the subnational level: An analysis of three countries. Habitat International, 33(3), 253-259. De Souza, A.P., M. Gaspar, E.A.d. Silva, E.C. Ulian, A.J. Waclawovsky, M.Y. Nishiyama Jr, R.V.d. Santos, M.M. Teixeira, G.M. Souza, and M.S. Buckeridge, 2008: Elevated CO(2) increases photosynthesis, biomass and productivity, and modifies gene expression in sugarcane. Plant Cell and Environment, 31(8), 1116-1127. De Souza, A.P., A. Grandis, D.C.C. Leite, and M.S. Buckeridge, 2013: Sugarcane as a Bioenergy Source: History, Performance, and Perspectives for Second-Generation Bioethanol. Bioenerg. Res., (published online). De Souza, H.N., R.G.M. De Goede, L. Brussaard, I.M. Cardoso, E.M.G. Duarte, R.B.A. Fernandes, L.C. Gomes, and M.M. Pulleman, 2012: Protective shade, tree diversity and soil properties in coffee agroforestry systems in the Atlantic Rainforest biome. Agriculture Ecosystems & Environment, 146(1), 179-196. Dearing, M.D. and L. Dizney, 2010: Ecology of hantavirus in a changing world. Annals of the New York Academy of Sciences, 1195(1), 99-112. Debels, P., C. Szlafsztein, P. Aldunce, C. Neri, Y. Carvajal, M. Quintero-Angel, A. Celis, A. Bezanilla, and D. Martínez, 2009: IUPA: a tool for the evaluation of the general usefulness of practices for adaptation to climate change and variability. Natural Hazards, 50(2), 211-233. Degallier, N., C. Favier, C. Menkes, M. Lengaigne, W.M. Ramalho, R. Souza, J. Servain, and J.-. Boulanger, 2010: Toward an early warning system for dengue prevention: Modeling climate impact on dengue transmission. Climatic Change, 98(3), 581-592. Demaria, E.M.C., E.P. Maurer, B. Thrasher, S. Vicuna, and F.J. Meza, 2013: Climate change impacts on an alpine watershed in Chile: do new model projections change the story? Journal of Hydrology, 502, 128-138. Dias, M.O.S., T.L. Junqueira, O. Cavalett, M.P. Cunha, C.D.F. Jesus, C.E.V. Rossell, R. Maciel Filho, and A. Bonomi, 2012: Integrated versus stand-alone second generation ethanol production from sugarcane bagasse and trash. Bioresource Technology, 103(1), 152-161. Diez Roux, A.V., T. Green Franklin, M. Alazraqui, and H. Spinelli, 2007: Intraurban variations in adult mortality in a large Latin American city. Journal of Urban Health, 84(3), 319-333. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Diffenbaugh, N.S., F. Giorgi, and J.S. Pal, 2008: Climate change hotspots in the United States. Geophysical Research Letters, 35(16), L16709. Diffenbaugh, N.S. and F. Giorgi, 2012: Climate change hotspots in the CMIP5 global climate model ensemble. Climatic Change, 114(3-4), 813-822. Döll, P., 2009: Vulnerability to the impact of climate change on renewable groundwater resources: a global-scale assessment. Environmental Research Letters, 4(3), 035006. Donat, M.G., L.V. Alexander, H. Yang, I. Durre, R. Vose, R.J.H. Dunn, K.M. Willett, E. Aguilar, M. Brunet, J. Caesar, B. Hewitson, C. Jack, A.M.G. Klein Tank, A.C. Kruger, J.A. Marengo, T.C. Peterson, M. Renom, C. Oria Rojas, M. Rusticucci, J. Salinger, A. Sanhouri Elrayah, S.S. Sekele, A.K. Srivastava, B. Trewin, C. Villarroel, L.A. Vincent, P. Zhai, X. Zhang, and S. Kitching, 2013: Updated analyses of temperature and precipitation extreme indices since the beginning of the twentieth century: The HadEX2 dataset. Journal of Geophysical Research: Atmospheres, 118(5), 2098-2118. Dos Santos, W.D., E.O. Gomez, and M.S. Buckeridge, 2011: Bioenergy and the Sustainable Revolution. In: Routes to Cellulosic Ethanol. [Buckeridge, M.S. and G.H. Goldmann (eds.)]. Springer, New York, USA, pp. 15-26. Doyle, M.E. and V.R. Barros, 2011: Attribution of the river flow growth in the Plata Basin. International Journal of Climatology, 31(15), 2234-2248. Doyle, M.E., R.I. Saurral, and V.R. Barros, 2012: Trends in the distributions of aggregated monthly precipitation over the La Plata Basin. International Journal of Climatology, 32(14), 2149-2162. Dufek, A.S. and T. Ambrizzi, 2008: Precipitation variability in Sao Paulo State, Brazil. Theoretical and Applied Climatology, 93(3-4), 167-178. Dufek, A.S., T. Ambrizzi, and R.P. da Rocha, 2008: Are Reanalysis Data Useful for Calculating Climate Indices over South America? In: Trends and Directions in Climate Research. Vol. 1146, Annals of the New York Academy of Sciences. [Gimeno, L., R. Garcia Herrera, and R.M. Trigo (eds.)]. Blackwell Publishing, New York, NY, USA, pp. 87-104. Duke, N.C., J.-. Meynecke, S. Dittmann, A.M. Ellison, K. Anger, U. Berger, S. Cannicci, K. Diele, K.C. Ewel, C.D. Field, N. Koedam, S.Y. Lee, C. Marchand, I. Nordhaus, and F. Dahdouh-Guebas, 2007: A World Without Mangroves? Science, 317(5834), 41-42. Dupnik, K.M., E.L. Nascimento, J.F. Rodrigues-Neto, T. Keesen, M. Zélia Fernandes, I. Duarte, and S.M.B. Jeronimo, 2011: New challenges in the epidemiology and treatment of visceral leishmaniasis in periurban areas. Drug Development Research, 72(6), 451-462. Dussaillant, A., G. Benito, W. Buytaert, P. Carling, C. Meier, and F. Espinoza, 2010: Repeated glacial-lake outburst floods in Patagonia: an increasing hazard? Natural Hazards, 54(2), 469-481. Eakin, C.M., J.A. Morgan, S.F. Heron, T.B. Smith, G. Liu, L. Alvarez-Filip, B. Baca, E. Bartels, C. Bastidas, C. Bouchon, M. Brandt, A.W. Bruckner, L. Bunkley-Williams, A. Cameron, B.D. Causey, M. Chiappone, T.R.L. Christensen, M.J.C. Crabbe, O. Day, E. de la Guardia, G. Diaz-Pulido, D. DiResta, D.L. Gil-Agudelo, D.S. Gilliam, R.N. Ginsburg, S. Gore, H.M. Guzman, J.C. Hendee, E.A. Hernandez-Delgado, E. Husain, C.F.G. Jeffrey, R.J. Jones, E. Jordan-Dahlgren, L.S. Kaufman, D.I. Kline, P.A. Kramer, J.C. Lang, D. Lirman, J. Mallela, C. Manfrino, J. Marechal, K. Marks, J. Mihaly, W.J. Miller, E.M. Mueller, E.M. Muller, C.A. Orozco Toro, H.A. Oxenford, D. Ponce-Taylor, N. Quinn, K.B. Ritchie, S. Rodriguez, A. Rodriguez Ramirez, S. Romano, J.F. Samhouri, J.A. Sanchez, G.P. Schmahl, B.V. Shank, W.J. Skirving, S.C.C. Steiner, E. Villamizar, S.M. Walsh, C. Walter, E. Weil, E.H. Williams, K.W. Roberson, and Y. Yusuf, 2010: Caribbean Corals in Crisis: Record Thermal Stress, Bleaching, and Mortality in 2005. Plos One, 5(11), e13969. Eakin, H. and M.C. Lemos, 2006: Adaptation and the state: Latin America and the challenge of capacity-building under globalization. Global Environmental Change, 16(1), 7-18. Eakin, H., L.A. Bojórquez-Tapia, R. Monterde Diaz, E. Castellanos, and J. Haggar, 2011: Adaptive Capacity and Social-Environmental Change: Theoretical and Operational Modeling of Smallholder Coffee Systems Response in Mesoamerican Pacific Rim. Environmental Management, 47(3), 352-367. Eakin, H., C. Tucker, E. Castellanos, R. Diaz-Porras, J. Barrera, and H. Morales, 2013: Adaptation in a multi- stressor environment: perceptions and responses to climatic and economic risks by coffee growers in Mesoamerica. Environment, Development and Sustainability, published online, 1-17. Eakin, H.C. and M.B. Wehbe, 2009: Linking local vulnerability to system sustainability in a resilience framework: two cases from Latin America. Climatic Change, 93(3-4), 355-377. Echeverría, C., D. Coomes, J. Salas, J.M. Rey-Benayas, A. Lara, and A. Newton, 2006: Rapid deforestation and fragmentation of Chilean Temperate Forests. Biological Conservation, 130(4), 481-494. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 ECLAC, 2008: Structural Change and Productivity Growth - 20 Years Later. Old problems, new opportunities. LC/G.2367(SES.32/3). Economic Commission for Latin America and the Caribbean (ECLAC), Santiago de Chile, Chile, pp. 328. ECLAC, 2009a: La Economia del Cambio Climatico en Chile: Sintesis. (LC/W.288). Economic Commission for Latin America and the Caribbean (ECLAC), Santiago de Chile, pp. 88. ECLAC, 2009b: Social Panorama of Latin America 2009. Briefing paper. United Nations, Santiago, Chile, pp. 64. ECLAC, 2009c: Cambio climático y desarrollo en América Latina y el Caribe: Una resena (LC/W.232.). Economic Commission for Latin America and the Caribbean (ECLAC), Santiago de Chile. ECLAC, 2010a: Economics of Climate Change in Latin America and the Caribbean. Summary 2010. (LC/G.2474). United Nations, ECLAC, Santiago de Chile, Chile, pp. 107. ECLAC, 2010b: El progreso de América Latina y el Caribe hacia los Objetivos de Desarrollo del Milenio. Desafíos para lograrlos con igualdad. (LC/G 2460). Available at: http://www.eclac.org/publicaciones/xml/1/39991/portada-indice-intro.pdf. Economic Commission for Latin America and the Caribbean (ECLAC), Santiago de Chile, Chile, pp. 25. ECLAC, 2010c: The Economics of Climate Change in Central America: Summary 2010. (LC/MEX/L.978). United Nations, ECLAC, pp. 146. ECLAC, 2010d: Economics of Climate Change in Latin America and the Caribbean. Summary 2010. United Nations, Economic Commission for Latin America and the Caribbean (ECLAC), Santiago, Chile, pp. 107. ECLAC, 2010e: Latin America and the Caribbean in the world economy. 2009-2010. A crisis generated in the centre and a recovery driven by the emerging economies . United Nations, Santiago, Chile, pp. 164. ECLAC, 2010f: The reactions of the Governments of the Americas to the international crisis: an overview of policy measures up to 31 December 2009. United Nations, ECLAC, Santiago, Chile, pp. 69. ECLAC, 2010g: Time for equality. Closing gaps, opening trails. In: Thirty-third session of ECLAC. Brasilia, 30 May to 1 June 2010. United Nations (UN), Santiago, Chile, pp. 269. ECLAC, 2011a: Efectos del cambio climático en la costa de América Latina y el Caribe : Dinámicas, tendencias y variabilidad climática. [Effects of climate change on the coast of Latin America and the Caribbean: dynamics, trends and climate variability.] LC/W.447. United Nations, Economic Commission for Latin America and the Caribbean (ECLAC), Santiago de Chile, Chile, pp. 263. ECLAC, 2011b: Social Panorama of Latin America 2010. United Nations, Santiago, Chile, pp. 252. ECLAC, 2011c: Social panorama of Latin America 2011. Briefing Paper. United Nations (UN), Santiago, Chile, pp. 49. ECLAC, 2011d: Subregión norte de América Latina y el Caribe: Información del sector agropecuario: las tendencias alimentarias 2000-2010. United Nations, México, D. F. ECLAC, 2012: Sustainable development 20 years on from the Earth Summit Summary. Progress, gaps and strategic guidelines for Latin America and the Caribbean. Summary. United Nations (UN), Santiago, Chile, pp. 55. Engel, S., S. Pagiola, and S. Wunder, 2008: Designing payments for environmental services in theory and practice: An overview of the issues. Ecological Economics, 65(4), 663-674. Engle, N.L., O.R. Johns, M.C. Lemos, and D.R. Nelson, 2011: Integrated and Adaptive Management of Water Resources: Tensions, Legacies, and the Next Best Thing. Ecology and Society, 16(1), 19. Engle, N.L. and M.C. Lemos, 2010: Unpacking governance: Building adaptive capacity to climate change of river basins in Brazil. Global Environmental Change-Human and Policy Dimensions, 20(1), 4-13. Englehart, P.J. and A.V. Douglas, 2006: Defining intraseasonal rainfall variability within the North American monsoon. Journal of Climate, 19(17), 4243-4253. Espinoza, J.C., J. Ronchail, F. Frappart, W. Lavado, W. Santini, and J.L. Guyot, 2013: The Major Floods in the Amazonas River and Tributaries (Western Amazon Basin) during the 1970-2012 Period: A Focus on the 2012 Flood. Journal of Hydrometeorology, 14(3), 1000-1008. Espinoza, J.C., J.L. Guyot, J. Ronchail, G. Cochonneau, N. Filizola, P. Fraizy, D. Labat, E. de Oliveira, J. Julio Ordonez, and P. Vauchel, 2009a: Contrasting regional discharge evolutions in the Amazon basin (1974-2004). Journal of Hydrology, 375(3-4), 297-311. Espinoza, J.C., J. Ronchail, J.L. Guyot, G. Cochonneau, F. Naziano, W. Lavado, E. De Oliveira, R. Pombosa, and P. Vauchel, 2009b: Spatio-temporal rainfall variability in the Amazon basin countries (Brazil, Peru, Bolivia, Colombia, and Ecuador). International Journal of Climatology, 29(11), 1574-1594. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Espinoza, J.C., J. Ronchail, J.L. Guyot, C. Junquas, P. Vauchel, W. Lavado, G. Drapeau, and R. Pombosa, 2011: Climate variability and extreme drought in the upper Solimoes River (western Amazon Basin): Understanding the exceptional 2010 drought. Geophysical Research Letters, 38(13), L13406. Espinoza, J.C., M. Lengaigne, J. Ronchail, and S. Janicot, 2012: Large-scale circulation patterns and related rainfall in the Amazon Basin: a neuronal networks approach. Climate Dynamics, 38(1-2), 121-140. Fábrega, J., T. Nakaegawa, R. Pinzón, K. Nakayama, O. Arakawa, and SOUSEI Theme-C Modeling Group, 2013: Hydroclimate projections for Panama in the late 21st Century. Hydrological Research Letters, 7(2), 23-29. Falvey, M. and R.D. Garreaud, 2009: Regional cooling in a warming world: Recent temperature trends in the southeast Pacific and along the west coast of subtropical South America (1979 2006). Journal of Geophysical Research, 114, D04102. FAO, 2009: Global Forest Resources Assessment 2010. Brazil Country Report. United Nations, Food and Agriculture Organization (FAO), Rome, pp. 111. FAO, 2010: FAO Forestry Paper 163. In: Global Forest Resources Assessment 2010 United Nations, Food and Agriculture Organization (FAO), Rome, pp. 340. FAO, 2012: The State of Food Insecurity in the World 2012. Economic growth is necessary but not sufficient to accelerate reduction of hunger and malnutrition. FAO, WFP and IFAD;, Rome, Italy. Farley, K.A., G. Pineiro, S.M. Palmer, E.G. Jobbágy, and R.B. Jackson, 2009: Stream acidification and base cation losses with grassland afforestation. Water Resources Research, 45(7). Fearnside, P.M., 2008: The Roles and Movements of Actors in the Deforestation of Brazilian Amazonia. Ecology and Society, 13(1), 23. Fearnside, P.M. and S. Pueyo, 2012: Greenhouse-gas emissions from tropical dams. Nature Climate Change, 2(6), 382-384. Feeley, K.J. and M.R. Silman, 2009: Extinction risks of Amazonian plant species. Proceedings of the National Academy of Sciences of the United States of America, 106(30), 12382-12387. Feliciangeli, M.D., O. Delgado, B. Suarez, and A. Bravo, 2006: Leishmania and sand flies: proximity to woodland as a risk factor for infection in a rural focus of visceral leishmaniasis in west central Venezuela; Leishmania et phlébotomes: La proximité des bois comme facteur de risque pour l'infection dans un foyer rural de leishmaniose viscérale dans le centre ouest du Venezuela; Leishmania y flebótomos: la proximidad al bosque como factor de riesgo de infección en un foco rural de leishmaniasis visceral en el centro-oeste de Venezuela. Tropical Medicine & International Health, 11(12), 1785-1791. Fernández, M.S., E.A. Lestani, R. Cavia, and O.D. Salomón, 2012: Phlebotominae fauna in a recent deforested area with American Tegumentary Leishmaniasis transmission (Puerto Iguazú, Misiones, Argentina): Seasonal distribution in domestic and peridomestic environments. Acta Tropica, 122(1), 16-23. Ferreira, A.R. and R.S.V. Teegavarapu, 2012: Optimal and Adaptive Operation of a Hydropower System with Unit Commitment and Water Quality Constraints. Water Resources Management, 26(3), 707-732. FEWS NET, 2013: Special Report Central America. Coffee sector shocks and projected food security impacts in Central America. Famine Early Warning Systems Network (FEWS NET), Washington, pp. 3pp. Ficke, A.D., C.A. Myrick, and L.J. Hansen, 2007: Potential impacts of global climate change on freshwater fisheries. Reviews in Fish Biology and Fisheries, 17(4), 581-613. Fiebig-Wittmaack, M., O. Astudillo, E. Wheaton, V. Wittrock, C. Perez, and A. Ibacache, 2012: Climatic trends and impact of climate change on agriculture in an arid Andean valley. Climatic Change, 111(3-4), 819-833. Finer, M. and C.N. Jenkins, 2012: Proliferation of Hydroelectric Dams in the Andean Amazon and Implications for Andes-Amazon Connectivity. Plos One, 7(4), e35126. FIOCRUZ, 2011: Mapa de Vulnerabilidade da Populaçao do Estado do Rio de Janeiro aos Impactos das Mudanças Climáticas nas Áreas Social, Saúde e Ambiente. Relatório 4, Versao final. FIOCRUZ, pp. 162. Fischedick, M., R. Schaeffer, A. Adedoyin, M. Akai, T. Bruckner, L. Clarke, V. Krey, I. Savolainen, S. Teske, D. Ürge-Vorsatz, and R. Wright, 2011: Mitigation Potential and Costs. In: IPCC Special Report on Renewable Energy Sources and Climate change Mitigation. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Fitzherbert, E., M. Struebig, A. Morel, F. Danielsen, C. Brühl, P. Donald, and B. Phalan, 2008: How will oil palm expansion affect biodiversity? Trends in Ecology and Evolution, 23(10), 538-545. Folke, C., S. Carpenter, T. Elmqvist, L. Gunderson, C.S. Holling, and B. Walker, 2002: Resilience and sustainable development: Building adaptive capacity in a world of transformations. Ambio, 31(5), 437-440. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Fortner, S.K., B.G. Mark, J.M. McKenzie, J. Bury, A. Trierweiler, M. Baraer, P.J. Burns, and L. Munk, 2011: Elevated stream trace and minor element concentrations in the foreland of receding tropical glaciers. Applied Geochemistry, 26(11), 1792-1801. Foster, J.L., D.K. Hall, R.E.J. Kelly, and L. Chiu, 2009: Seasonal snow extent and snow mass in South America using SMMR and SSM/I passive microwave data (1979 2006). Remote Sensing of Environment, 113(2), 291- 305. Francini-Filho, R.B. and R.L. Moura, 2008: Dynamics of fish assemblages on coral reefs subjected to different management regimes in the Abrolhos Bank, eastern Brazil. Aquatic Conservation-Marine and Freshwater Ecosystems, 18(7), 1166-1179. Francini-Filho, R.B., R.L. Moura, F.L. Thompson, R.M. Reis, L. Kaufman, R.K.P. Kikuchi, and Z.M. Leao, 2008: Diseases leading to accelerated decline of reef corals in the largest South Atlantic reef complex (Abrolhos Bank, eastern Brazil). Marine Pollution Bulletin, 56(5), 1008-1014. Franco-Paredes, C., D. Jones, A.J. Rodriguez-Morales, and J. Ignacio Santos-Preciado, 2007: Commentary: improving the health of neglected populations in Latin America. Bmc Public Health, 7, 11. Francou, B., B. Cáceres, J. Gomez, and A. Soruco, 2007: Coherence of the glacier signal throughout the tropical Andes over the last decades. Proceedings of Proceedings of the First International Conference on the Impact of Climate Change on High-Mountain System. IDEAM , 5, November 2005, Bogota D.C., Colombia, pp. 87-97. Francou, B., D. Fabre, B. Pouyaud, V. Jomelli, and Y. Arnaud, 1999: Symptoms of degradation in a tropical rock glacier, Bolivian Andes. Permafrost and Periglacial Processes, 10(1), 91-100. Fraser, B., 2012: Melting in the Andes: Goodbye glaciers. Nature, 491(7423), 180 182. Freire, K.M.F. and D. Pauly, 2010: Fishing down Brazilian marine food webs, with emphasis on the east Brazil large marine ecosystem. Fisheries Research, 105(1), 57-62. Freitas, M.A.V. and J.L.S. Soito, 2009: Vulnerability to climate change and water management: hydropower generation in Brazil. River Basin Management V, 124, 217-226. Fry, L., D. Watkins, J. Mihelcic, and N. Reents, 2010: Sustainability of Gravity Fed Water Systems in Alto Beni, Bolivia: Preparing for Change[Palmer, R.N. (ed.)]. Proceedings of World Environmental and Water Resources Congress 2010: Challenges of Change, Providence, Rhode Island, USA, 16-20 May, 2010, pp. 751. Fry, L.M., D.W. Watkins, N. Reents, M.D. Rowe, and J.R. Mihelcic, 2012: Climate change and development impacts on the sustainability of spring-fed water supply systems in the Alto Beni region of Bolivia. Journal of Hydrology, 468, 120-129. Fuentes, M.V., 2004: Proposal of a Geographic Information System for modeling zoonotic fasciolosis transmission in the Andes. Parasitologia Latinoamericana, 59(1-2), 51-55. Fuller, D.O., A. Troyo, and J.C. Beier, 2009: El Nino Southern Oscillation and vegetation dynamics as predictors of dengue fever cases in Costa Rica. Environmental Research Letters, 4(1). García, A.L., R. Parrado, E. Rojas, R. Delgado, J.-. Dujardin, and R. Reithinger, 2009: Leishmaniases in bolivia: Comprehensive review and current status. American Journal of Tropical Medicine and Hygiene, 80(5), 704- 711. Gardner, C.L. and K.D. Ryman, 2010: Yellow fever: A reemerging threat. Clinics in Laboratory Medicine, 30(1), 237-260. Garreaud, R.D. and M. Falvey, 2009: The coastal winds off western subtropical South America in future climate scenarios. International Journal of Climatology, 29(4), 543-554. Gascoin, S., C. Kinnard, R. Ponce, S. Lhermitte, S. MacDonell, and A. Rabatel, 2011: Glacier contribution to streamflow in two headwaters of the Huasco River, Dry Andes of Chile. The Cryosphere, 5(4), 1099-1113. Gasparri, N.I., H.R. Grau, and E. Manghi, 2008: Carbon pools and emissions from deforestation in extra-tropical forests of Northern Argentina between 1900 and 2005. Ecosystems, 11, 1247-1261. Gasparri, N.I. and H.R. Grau, 2009: Deforestation and fragmentation of Chaco dry forest in NW Argentina (1972- 2007). Forest Ecology and Management, 258(6), 913-921. Gasper, R., A. Blohm, and M. Ruth, 2011: Social and economic impacts of climate change on the urban environment. Current Opinion in Environmental Sustainability, 3(3), 150-157. Gavilán, R.G. and J. Martínez-Urtaza, 2011: Environmental drivers of emergence and spreading of vibrio epidemics in South America. Revista Peruana De Medicina De Experimental y Salud Publica, 28(1), 109-115. Geerts, S., D. Raes, and M. Garcia, 2010: Using AquaCrop to derive deficit irrigation schedules. Agricultural Water Management, 98(1), 213-216. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Geerts, S. and D. Raes, 2009: Deficit irrigation as an on-farm strategy to maximize crop water productivity in dry areas. Agricultural Water Management, 96(9), 1275-1284. Georgescu, M., D.B. Lobell, C.B. Field, and A. Mahalov, 2013: Simulated hydroclimatic impacts of projected Brazilian sugarcane expansion. Geophysical Research Letters, 40(5), 972-977. Gharbi, M., P. Quenel, J. Gustave, S. Cassadou, G. La Ruche, L. Girdary, and L. Marrama, 2011: Time series analysis of dengue incidence in Guadeloupe, French West Indies: Forecasting models using climate variables as predictors. Bmc Infectious Diseases, 11, 166. Ghini, R., W. Bettiol, and E. Hamada, 2011: Diseases in tropical and plantation crops as affected by climate changes: current knowledge and perspectives. Plant Pathology, 60(1), 122-132. Gibbons, J.M., E. Nicholson, E.J. Milner-Gulland, and J.P.G. Jones, 2011: Should payments for biodiversity conservation be based on action or results? Journal of Applied Ecology, 48(5), 1218-1226. Gil, L.H.S., M.S. Tada, T.H. Katsuragawa, P.E.M. Ribolla, and L.H.P. Da Silva, 2007: Urban and suburban malaria in Rondônia (Brazilian Western Amazon) II. Perennial transmissions with high anopheline densities are associated with human environmental changes. Memorias do Instituto Oswaldo Cruz, 102(3), 271-276. Gilbert, A., P. Wagnon, C. Vincent, P. Ginot, and M. Funk, 2010: Atmospheric warming at a high-elevation tropical site revealed by englacial temperatures at Illimani, Bolivia (6340 m above sea level, 16°S, 67°W). Journal of Geophysical Research, 115(D10). Gilman, E.L., J. Ellison, N.C. Duke, and C. Field, 2008: Threats to mangroves from climate change and adaptation options: A review. Aquatic Botany, 89(2), 237-250. Giorgi, F., 2002: Variability and trends of sub-continental scale surface climate in the twentieth century. Part I: observations. Climate Dynamics, 18(8), 675-691. Giorgi, F., 2006: Climate change hot-spots. Geophysical Research Letters, 33(8), L08707. Giorgi, F. and N. Diffenbaugh, 2008: Developing regional climate change scenarios for use in assessment of effects on human health and disease. Climate Research, 36(2), 141-151. Giraldo, D.H.J., W. Pérez, I. Trebejo, W. Yzarra, and G. Forbes, 2010: Severidad del tizón tardío de la papa (Phytophthora infestans) en zonas agrícolas del Perú asociado con el cambio climático. Revista Peruana Geo- Atmosférica (RPGA), 2, 56-67. Giri, C., E. Ochieng, L.L. Tieszen, Z. Zhu, A. Singh, T. Loveland, J. Masek, and N. Duke, 2011: Status and distribution of mangrove forests of the world using earth observation satellite data. Global Ecology and Biogeography, 20(1), 154-159. Gloor, M., R.J.W. Brienen, D. Galbraith, T.R. Feldpausch, J. Schöngart, J.-. Guyot, J.C. Espinoza, J. Lloyd, and O.L. Phillips, 2013: Intensification of the Amazon hydrological cycle over the last two decades. Geophysical Research Letters, 40(9), 1729-1733. Goldemberg, J., 2008: The Brazilian biofuels industry. Biotechnology for Biofuels, 1(1), 6. Gomes, A.F., A.A. Nobre, and O.G. Cruz, 2012: Temporal analysis of the relationship between dengue and meteorological variables in the city of Rio de Janeiro, Brazil, 2001-2009. Cad. Saúde Pública [Online], 28(11), 2189-2197. Gomez, C., A.J. Rodriguez-Morales, and C. Franco-Paredes, 2006: Impact of climate variability in the occurrence of leishmaniasis in Bolivia. American Journal of Tropical Medicine and Hygiene, 75(5), 42-42. Gondim, R.S., M.A.H. de Castro, A.d.H.N. Maia, S.R.M. Evangelista, and Fuck, Sergio C. de F.,Jr., 2012: Climate Change Impacts on Irrigation Water Needs in the jaguaribe River Basin. Journal of the American Water Resources Association, 48(2), 355-365. Gondim, R.S., Holanda de Castro, Marco Aurélio Medeiros Evangelista,Sílvio Roberto de, A.d. Santos Teixeira, and França Fuck Júnior,Sérgio César de, 2008: Climate change and impacts on water requirement of permanent crops in the Jaguaribe Basin, Ceará, Brazil. Pesquisa Agropecuária Brasileira, 43(12). Gosling, S.N., R.G. Taylor, N.W. Arnell, and M.C. Todd, 2011: A comparative analysis of projected impacts of climate change on river runoff from global and catchment-scale hydrological models. Hydrology and Earth System Sciences, 15(1), 279-294. Gottdenker, N.L., J.E. Calzada, A. Saldana, and C.R. Carroll, 2011: Association of anthropogenic land use change and increased abundance of the Chagas disease vector Rhodnius pallescens in a rural landscape of Panama. American Journal of Tropical Medicine and Hygiene, 84(1), 70-77. Graham, C., L. Higuera, and E. Lora, 2011: Which health conditions cause the most unhappiness? Health Economics, 20(12), 1431-1447. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Grantz, K., B. Rajagopalan, M. Clark, and E. Zagona, 2007: Seasonal Shifts in the North American Monsoon. Journal of Climate, 20(9), 1923-1935. Grass, D. and M. Cane, 2008: The effects of weather and air pollution on cardiovascular and respiratory mortality in Santiago, Chile, during the winters of 1988-1996. International Journal of Climatology, 28(8), 1113-1126. Gratiot, N., E.J. Anthony, A. Gardel, C. Gaucherel, C. Proisy, and J.T. Wells, 2008: Significant contribution of the 18.6 year tidal cycle to regional coastal changes. Nature Geoscience, 1(3), 169-172. Grau, H.R. and M. Aide, 2008: Globalization and land-use transitions in Latin America. Ecology and Society, 13(2), 16. Gray, C.L., R.E. Bilsborrow, J.L. Bremner, and F. Lu, 2008: Indigenous land use in the Ecuadorian Amazon: A cross-cultural and multilevel analysis. Human Ecology, 36(1), 97-109. Gregg, J.S. and S.J. Smith, 2010: Global and regional potential for bioenergy from agricultural and forestry residue biomass. Mitigation and Adaptation Strategies for Global Change, 15(3), 241-262. Gruskin, D., 2012: Agbiotech 2.0. Nature Biotechnology, 30(3), 211-214. Guarderas, A.P., S.D. Hacker, and J. Lubchenco, 2008: Current Status of Marine Protected Areas in Latin America and the Caribbean. Conservation Biology, 22(6), 1630-1640. Guariguata, M.R., 2009: Tropical Forest Management and Climate Change Adaptation. Revista De Estudios Sociales,(32), 98-112. Guariguata, M.R., P. Sist, and R. Nasi, 2012: Reprint of: multiple use management of tropical production forests: how can we move from concept to reality? Forest Ecology and Management, 268, 1-5. Guevara, S. and J. Laborde, 2008: The Landscape Approach: Designing New Reserves for Protection of Biological and Cultural Diversity in Latin America. Environmental Ethics, 30(3), 251-262. Guimberteau, M., J. Ronchail, J.C. Espinoza, M. Lengaigne, B. Sultan, J. Polcher, G. Drapeau, J.-. Guyot, A. Ducharne, and P. Ciais, 2013: Future changes in precipitation and impacts on extreme streamflow over Amazonian sub-basins. Environmental Research Letters, 8(1), 014035. Gullison, R.E., P.C. Frumhoff, J.G. Canadell, C.B. Field, D.C. Nepstad, K. Hayhoe, R. Avissar, L.M. Curran, P. Friedlingstein, C.D. Jones, and C. Nobre, 2007: Tropical Forests and Climate Policy. Science, 316(5827), 985- 986. Gurjar, B.R., A. Jain, A. Sharma, A. Agarwal, P. Gupta, A.S. Nagpure, and J. Lelieveld, 2010: Human health risks in megacities due to air pollution. Atmospheric Environment, 44(36), 4606-4613. Gutiérrez, D., A. Bertrand, C. Wosnitza-mendo, B. Dewitte, S. Purca, C. Pena, A. Chaigneau, J. Tam, M. Graco, C. Grados, P. Fréon, and R. Guevara-carrasco, 2011a: Sensibilidad del sistema de afloramiento costero del Perú al cambio climático e implicancias ecológicas [Climate change sensitivity of the Peruvian upwelling system and ecological implications]. Revista Peruana Geoatmosférica, 3, 1-24. Gutiérrez, D., I. Bouloubassi, A. Sifeddine, S. Purca, K. Goubanova, M. Graco, D. Field, L. Mejanelle, F. Velazco, A. Lorre, R. Salvatteci, D. Quispe, G. Vargas, B. Dewitte, and L. Ortlieb, 2011b: Coastal cooling and increased productivity in the main upwelling zone off Peru since the mid-twentieth century. Geophysical Research Letters, 38, L07603. Gutiérrez-Vélez, V., R.S. DeFries, M. Pinedo-Vásquez, M. Uriarte, C. Padoch, W.E. Baethgen, K. Fernandes, and Y. Lim, 2011: High-yield oil palm expansion spares land at the expense of forests in the Peruvian Amazon. Environmental Research Letters, 6(4), 044029. Hajat, S., M. O'Connor, and T. Kosatsky, 2010: Health effects of hot weather: from awareness of risk factors to effective health protection. The Lancet, 375(9717), 856-863. Hall, T.C., A.M. Sealy, T.S. Stephenson, S. Kusunoki, M.A. Taylor, A.A. Chen, and A. Kitoh, 2013: Future climate of the Caribbean from a super-high-resolution atmospheric general circulation model. Theoretical and Applied Climatology, 113(1-2), 271-287. Halpern, B.S., S. Walbridge, K.A. Selkoe, C.V. Kappel, F. Micheli, C. D'Agrosa, J.F. Bruno, K.S. Casey, C. Ebert, H.E. Fox, R. Fujita, D. Heinemann, H.S. Lenihan, E.M.P. Madin, M.T. Perry, E.R. Selig, M. Spalding, R. Steneck, and R. Watson, 2008: A Global Map of Human Impact on Marine Ecosystems. Science, 319(5865), 948-952. Halpern, B.S., C. Longo, D. Hardy, K.L. McLeod, J.F. Samhouri, S.K. Katona, K. Kleisner, S.E. Lester, J. O'Leary, M. Ranelletti, A.A. Rosenberg, C. Scarborough, E.R. Selig, B.D. Best, D.R. Brumbaugh, F.S. Chapin, L.B. Crowder, K.L. Daly, S.C. Doney, C. Elfes, M.J. Fogarty, S.D. Gaines, K.I. Jacobsen, L.B. Karrer, H.M. Leslie, E. Neeley, D. Pauly, S. Polasky, B. Ris, K. St Martin, G.S. Stone, U.R. Sumaila, and D. Zeller, 2012: An index to assess the health and benefits of the global ocean. Nature, 488(7413), 615-620. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Halsnaes, K. and J. Verhagen, 2007: Development based climate change adaptation and mitigation conceptual issues and lessons learned in studies in developing countries. Mitigation and Adaptation Strategies for Global Change, 12(5), 665-684. Hanf, M., A. Adenis, M. Nacher, and B. Carme, 2011: The role of El Nino southern oscillation (ENSO) on variations of monthly Plasmodium falciparum malaria cases at the cayenne general hospital, 1996-2009, French Guiana. Malaria Journal, 10, 100. Hantke Domas, M., 2011: Avances legislativos en gestión sostenible y descentralizada del agua en América Latina.(LC/W.44). Economic Commission for Latin America and the Caribbean (ECLAC), Santiago de Chile, Chile, pp. 70. Hardoy, J. and G. Pandiella, 2009: Urban poverty and vulnerability to climate change in Latin America. Environment and Urbanization, 21(1), 203-224. Hardoy, J. and P. Romero-Lankao, 2011: Latin American cities and climate change: challenges and options to mitigation and adaptation responses. Current Opinion in Environmental Sustainability, 3(3), 158-163. Harrison, S., N. Glasser, V. Winchester, E. Haresign, C. Warren, and K. Jansson, 2006: A glacial lake outburst flood associated with recent mountain glacier retreat, Patagonian Andes. The Holocene, 16(4), 611-620. Harvey, C.A., O. Komar, R. Chazdon, B.G. Ferguson, B. Finegan, D.M. Griffith, M. Martínez-Ramos, H. Morales, R. Nigh, L. Soto-Pinto, M. Van Breugel, and M. Wishnie, 2008: Integrating agricultural landscapes with biodiversity conservation in the Mesoamerican hotspot. Conservation Biology, 22(1), 8-15. Hastenrath, S., 2012: Exploring the climate problems of Brazil's Nordeste: a review. Climatic Change, 112(2), 243- 251. Hastings, J.G., 2011: International Environmental NGOs and Conservation Science and Policy: A Case from Brazil. Coastal Management, 39(3), 317-335. Hayhoe, S.J., C. Neill, S. Porder, R. McHorney, P. Lefebvre, M.T. Coe, H. Elsenbeer, and A.V. Krusche, 2011: Conversion to soy on the Amazonian agricultural frontier increases streamflow without affecting stormflow dynamics. Global Change Biology, 17(5), 1821-1833. Hecht, S.B. and S.S. Saatchi, 2007: Globalization and forest resurgence: Changes in forest cover in El Salvador. Bioscience, 57(8), 663-672. Hegglin, E. and C. Huggel, 2008: An Integrated Assessment of Vulnerability to Glacial Hazards. A Case Study in the Cordillera Blanca, Peru. Mountain Research and Development, 28(3-4), 299-309. Heinrichs, D. and K. Krellenberg, 2011: Climate Adaptation Strategies: Evidence from Latin American City- Regions. Springer, New York, USA, pp. 230. Heller, N.E. and E.S. Zavaleta, 2009: Biodiversity management in the face of climate change: A review of 22 years of recommendations. Biological Conservation, 142(1), 14-32. Henríquez Ruiz, C., 2009: El proceso de urbanización en la cuenca del río Chillán y su capacidad adaptativa ante precipitaciones extremas [The process of urbanization in the Chillán s watershedChillán and its adaptive capacity to stormwater]. Revista Estudios Geográficos, 70(266), 155-179. Herrera-Martinez, A.D. and A.J. Rodríguez-Morales, 2010: Potential influence of climate variability on dengue incidence registered in a western pediatric Hospital of Venezuela. Tropical Biomedicine, 27(2), 280-286. Hertel, T.W., M.B. Burke, and D.B. Lobell, 2010: The poverty implications of climate-induced crop yield changes by 2030. Global Environmental Change, 20(4), 577-585. Hickey, C. and T. Weis, 2012: The challenge of climate change adaptation in Guyana. Climate and Development, 4(1), 66-74. Hidalgo, H.G., J.A. Amador, E.J. Alfaro, and B. Quesada, 2013: Hydrological climate change projections for Central America. Journal of Hydrology, 495, 94-112. Hirata, R. and B.P. Conicelli, 2012: Groundwater resources in Brazil: a review of possible impacts caused by climate change. An Acad Bras Cienc., 84(2), 297-312. Hoegh-Guldberg, O. and J.F. Bruno, 2010: The Impact of Climate Change on the World s Marine Ecosystems. Science, 328(5985), 1523-1528. Hoffmann, D., I. Oetting, C.A. Arnillas, and R. Ulloa, 2011: Climate Change and Protected Areas in the Tropical Andes. In: Climate Change and Biodiversity in the Tropical Andes. [Herzog, S.K., R. Martínez, P.M. Jrgensen, and H. Tiessen (eds.)]. Inter-American Institute for Global Change Research (IAI) and Scientific Committee on Problems of the Envrionment (SCOPE)., , pp. 311-325. Hofstra, N., 2011: Quantifying the impact of climate change on enteric waterborne pathogen concentrations in surface water. Current Opinion in Environmental Sustainability, 3(6), 471-479. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Holder, C.D., 2006: The hydrological significance of cloud forests in the Sierra de las Minas Biosphere Reserve, Guatemala. Geoforum, 37(1), 82-93. Holmner, A., A. Mackenzie, and U. Krengel, 2010: Molecular basis of cholera blood-group dependence and implications for a world characterized by climate change. FEBS Letters, 584(12), 2548-2555. Holt-Gimenez, E., 2002: Measuring farmers' agroecological resistance after Hurricane Mitch in Nicaragua: a case study in participatory, sustainable land management impact monitoring. Agriculture Ecosystems & Environment, 93(1-3), 87-105. Honório, N.A., C.T. Codeço, F.C. Alves, M.A.F.M. Magalhes, and R. Lourenço-De-Oliveira, 2009: Temporal distribution of aedes aegypti in different districts of Rio De Janeiro, Brazil, measured by two types of traps. Journal of Medical Entomology, 46(5), 1001-1014. Hotez, P.J., M.E. Bottazzi, C. Franco-Paredes, S.K. Ault, and M.R. Periago, 2008: The neglected tropical diseases of Latin America and the Caribbean: A review of disease burden and distribution and a roadmap for control and elimination. PLoS Neglected Tropical Diseases, 2(9). Hoyos, L.E., A.M. Cingolani, M.R. Zak, M.V. Vaieretti, D.E. Gorla, and M.R. Cabido, 2012: Deforestation and precipitation patterns in the arid Chaco forests of central Argentina. Applied Vegetation Science, (published online 9 July 2012). Hoyos, N., J. Escobar, J.C. Restrepo, A.M. Arango, and J.C. Ortiz, 2013: Impact of the 2010-2011 La Nina phenomenon in Colombia, South America: The human toll of an extreme weather event. Applied Geography, 39, 16-25. Hsiang, S.M., 2010: Temperatures and cyclones strongly associated with economic production in the Caribbean and Central America. Proceedings of the National Academy of Sciences, USA, 107(35), 15367-15372. Huang, C., A.G. Barnett, X. Wang, P. Vaneckova, G. Fitzgerald, and S. Tong, 2011: Projecting Future heat-related mortality under climate change scenarios: A systematic review. Environmental Health Perspectives, 119(12), 1681-1690. Huarcaya, E., C. Maguina, R. Torres, J. Rupay, and L. Fuentes, 2004: Bartonelosis (Carrion's Disease) in the pediatric population of Peru: an overview and update. The Brazilian Journal of Infectious Diseases : An Official Publication of the Brazilian Society of Infectious Diseases, 8(5), 331-339. Hubbell, S.P., F. He, R. Condit, L. Borda-de-Agua, J. Kellner, and H. ter Steege, 2008: How many tree species and how many of them are there in the Amazon will go extinct? Proceedings of the National Academy of Sciences of the United States of America, 105, 11498-11504. ICO, 2013: Report on the outbreak of coffee leaf rust in Central America and Action Plan to combat the pest. (ED 2157/13). International Coffee Organization (ICO), London, UK, pp. 6pp. IDB, 2013: Emerging and Sustainable Cities Initiative (ESCI) (Accessible at: http://www.iadb.org/en/topics/emerging-and-sustainable-cities/emerging-and-sustainable-cities- initiative,6656.html) Inter-American Development Bank (IDB). IDEAM, 2012: Glaciares de Colombia, más que montanas con hielo. Instituto de Hidrología, Meteorología y Estudios Ambientales (IDEAM), Bogotá, D.C., Colombia, pp. 344. IEA, 2012: Statistics & Balances. In: IAE Statistics. Available at: http://www.iea.org/stats/index.asp. OECD, International Energy Agency (IEA), . Igreja, R.P., 2011: Infectious disease control in Brazil. The Lancet, 378(9797), 1135. Imbach, P., L. Molina, B. Locatelli, O. Roupsard, G. Mahé, R. Neilson, L. Corrales, M. Scholze, and P. Ciais, 2012: Modeling potential equilibrium states of vegetation and terrestrial water cycle of Mesoamerica under climate change scenarios. Journal of Hydrometeorology, 13(2), 665-680. INPE, 2011: Projeto PRODES. Monitoramente da Floresta Amazônica Brasileira por Satélite. Available at: http://www.obt.inpe.br/prodes/. Instituto Nacional de Pesquisas Espaciais (INPE), . Instituto Nacional de Estadística, 2011: Compendio Estadístico Ambiental de Guatemala 2010. Sección de Estadísticas Ambientales, Oficina Coordinadora Sectorial de Estadísticas de Ambiente y Recursos Naturales. OCSE/Ambiente, Guatemala, pp. 357. IPCC (ed.), 2011: IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. Prepared by Working Group III of the Intergovernmental Panel on Climate Change. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner, T. Zwickel, P. Eickemeier, G. Hansen, S. Schlömer, C. von Stechow (eds)] Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1075. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 IPCC, 2012a: Summary for Policymakers. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al. (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 1-19. IPCC, 2012b: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. In: A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change. [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al. (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 582. Ison, R., 2010: Systems Practice: How To Act In A Climate-change World. SPRINGER, London, UK, The Open University ed., pp. 324. ITC, 2007: Organic Farming and Climate Change. International Trade Centre UNCTAD/WTO; Research Institute of Organic Agriculture (FiBL), Geneva, Switzerland, pp. 27. Izaguirre, C., F. Méndez, A. Espejo, I.J. Losada, and B.G. Reguero, 2013: Extreme wave climate changes in Central-South America. Climatic Change, 119(2), 277-290. Izquierdo, A.E., C.D. De Angelo, and T.M. Aide, 2008: Thirty Years of Human Demography and Land-Use Change in the Atlantic Forest of Misiones, Argentina: an Evaluation of the Forest Transition Model. Ecology and Society, 13(2), 3. Jara-Rojas, R., B.E. Bravo-Ureta, and J. Díaz, 2012: Adoption of water conservation practices: A socioeconomic analysis of small-scale farmers in Central Chile. Agricultural Systems, 110, 54-62. Jarvis, A., J.L. Touval, M.C. Schmitz, L. Sotomayor, and G.G. Hyman, 2010: Assessment of threats to ecosystems in South America. Journal for Nature Conservation, 18(3), 180-188. Jasinski, R., L.A.A. Pereira, and A.L.F. Braga, 2011: Air pollution and pediatric hospital admissions due to respiratory diseases in Cubatao, Sao Paulo State, Brazil, from 1997 to 2004. Cadernos De Saude Publica, 27(11), 2242-2252. Jentes, E.S., G. Poumerol, M.D. Gershman, D.R. Hill, J. Lemarchand, R.F. Lewis, J.E. Staples, O. Tomori, A. Wilder-Smith, and T.P. Monath, 2011: The revised global yellow fever risk map and recommendations for vaccination, 2010: consensus of the Informal WHO Working Group on Geographic Risk for Yellow Fever. The Lancet Infectious Diseases, 11(8), 622-632. Jomelli, V., M. Khodri, V. Favier, D. Brunstein, M.P. Ledru, P. Wagnon, P.H. Blard, J.E. Sicart, R. Braucher, D. Grancher, D.L. Bourles, P. Braconnot, and M. Vuille, 2011: Irregular tropical glacier retreat over the Holocene epoch driven by progressive warming. Nature, 474(7350), 196-199. Jomelli, V., V. Favier, A. Rabatel, D. Brunstein, G. Hoffmann, and B. Francou, 2009: Fluctuations of glaciers in the tropical Andes over the last millennium and palaeoclimatic implications: A review. Palaeogeography, Palaeoclimatology, Palaeoecology, 281(3-4), 269-282. Jones, C. and L.V. Carvalho, 2013: Climate change in the South American Monsoon System: present climate and CMIP5 projections. Journal of Climate, 26(17), 6660-6678. Jonsson, C.B., L.T.M. Figueiredo, and O. Vapalahti, 2010: A global perspective on hantavirus ecology, epidemiology, and disease. Clinical Microbiology Reviews, 23(2), 412-441. Juen, I., G. Kaser, and C. Georges, 2007: Modelling observed and future runoff from a glacierized tropical catchment (Cordillera Blanca, Perú). Global and Planetary Change, 59(1-4), 37-48. Jutla, A.S., A.S. Akanda, and S. Islam, 2010: Tracking cholera in coastal regions using satellite observations. Journal of the American Water Resources Association, 46(4), 651-662. Kacef, O. and R. López-Monti, 2010: Latin America, from boom to crisis: macroeconomic policy challenges. Cepal Review, 100, 41-67. Kaimowitz, D., 2008: The prospects for Reduced Emissions from Deforestation and Degradation (REDD) in Mesoamerica. International Forestry Review, 10(3), 485-495. Kaimowitz, D. and A. Angelsen, 2008: Will livestock intensification help save Latin America's tropical forests? Journal of Sustainable Forestry, 27(1/2), 6-24. Kamiguchi, K., A. Kitoh, T. Uchiyama, R. Mizuta, and A. Noda, 2006: Changes in Precipitation-based Extremes Indices Due to Global Warming Projected by a Global 20-km-mesh Atmospheric Model. SOLA, 2, 64-67. Kane, E.M., R.M. Turcios, M.L. Arvay, S. Garcia, J.S. Bresee, and R.I. Glass, 2004: The epidemiology of rotavirus diarrhea in Latin America. Anticipating rotavirus vaccines. Revista Panamericana De Salud Publica/Pan American Journal of Public Health, 16(6), 371-377. Karanja, D., S.J. Elliott, and S. Gabizon, 2011: Community level research on water health and global change: where have we been? Where are we going? Current Opinion in Environmental Sustainability, 3(6), 467-470. Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Karmalkar, A.V., R.S. Bradley, and H.F. Diaz, 2011: Climate change in Central America and Mexico: regional climate model validation and climate change projections. Climate Dynamics, 37(3-4), 605-629. Kaser, G., M. Großhauser, and B. Marzeion, 2010: Contribution potential of glaciers to water availability in different climate regimes. Proceedings of the National Academy of Sciences, (published online before print). Keim, M.E., 2008: Building Human Resilience: The Role of Public Health Preparedness and Response As an Adaptation to Climate Change. American Journal of Preventive Medicine, 35(5), 508-516. Keller, M., D. Echeverría, and J. Parry, 2011a: Review of Current and Planned Adaptation Action: Central America and Mexico. Belize, Costa Rica, El Salvador, Guatemala, Honduras, Nicaragua, Mexico and Panama. International Institute for Sustainable Development (IISD), Adaptation Partnership, pp. 143pp. Keller, M., D. Medeiros, D. Echeverría, and J. Parry, 2011b: Review of Current and Planned Adaptation Action: South America. Argentina, Bolivia, Brazil, Chile, Colombia, Ecuador, Guyana, Paraguay, Peru, Suriname, Uruguay and Venezuela. International Institute for Sustainable Development (IISD), Adaptation Partnership, pp. 190. Kelly-Hope, L.A. and M. Thomson, 2010: Climate and Infectious Diseases (Chapter 4) . In: Seasonal Forecasts, Climate Change and Human Health. Springer Health and Climate Series: Advances in Global Change Research , Vol.30 [Thomson, M.C., R. Garcia-Herrera, and M. Beniston (eds.)]. Springer, London, UK, pp. 31-70. Kemkes, R.J., J. Farley, and C.J. Koliba, 2010: Determining when payments are an effective policy approach to ecosystem service provision. Ecological Economics, 69(11), 2069-2074. Khalil, A.F., H. Kwon, U. Lall, M.J. Miranda, and J. Skees, 2007: El Nino Southern Oscillation based index insurance for floods: Statistical risk analyses and application to Peru. Water Resources Research, 43(10). Killeen, T.J., A. Guerra, M. Calzada, L. Correa, V. Calderon, L. Soria, B. Quezada, and M.K. Steininger, 2008: Total Historical Land-Use Change in Eastern Bolivia: Who, Where, When, and How Much? Ecology and Society, 13(1), 36. Kirchhoff, C.J., M.C. Lemos, and N.L. Engle, 2013: What influences climate information use in water management? The role of boundary organizations and governance regimes in Brazil and the U.S. Environmental Science & Policy, 26, 6-18. Kitoh, A., H. Endo, K. Krishna Kumar, I.F.A. Cavalcanti, P. Goswami, and T. Zhou, 2013: Monsoons in a changing world: A regional perspective in a global context. Journal of Geophysical Research: Atmospheres, 118(8), 3053-3065. Kjellstrom, T., A.J. Butler, R.M. Lucas, and R. Bonita, 2010: Public health impact of global heating due to climate change: Potential effects on chronic non-communicable diseases. International Journal of Public Health, 55(2), 97-103. Kjellstrom, T. and J. Crowe, 2011: Climate change, workplace heat exposure, and occupational health and productivity in Central America. International Journal of Occupational and Environmental Health, 17(3), 270- 281. Kjellstrom, T., R.S. Kovats, S.J. Lloyd, T. Holt, and R.S.J. Tol, 2009: The Direct Impact of Climate Change on Regional Labor Productivity. Archives of Environmental & Occupational Health, 64(4), 217-227. Klemm, O., R.S. Schemenauer, A. Lummerich, P. Cereceda, V. Marzol, D. Corell, J. van Heerden, D. Reinhard, T. Gherezghiher, J. Olivier, P. Osses, J. Sarsour, E. Frost, M.J. Estrela, J.A. Valiente, and G.M. Fessehaye, 2012: Fog as a Fresh-Water Resource: Overview and Perspectives. AMBIO, 41, 221-234. Klimas, C.A., K.A. Kainer, and L.H.d.O. Wadt, 2012: The economic value of sustainable seed and timber harvests of multi-use species: An example using Carapa guianensis. Forest Ecology and Management, 268, 81-91. Koelle, K., 2009: The impact of climate on the disease dynamics of cholera. Clinical Microbiology and Infection, 15(SUPPL. 1), 29-31. Koh, L.P. and J. Ghazoul, 2008: Biofuels, biodiversity, and people: Understanding the conflicts and finding opportunities. Biological Conservation, 141(10), 2450-2460. Kosaka, Y. and S. Xie, 2013: Recent global-warming hiatus tied to equatorial Pacific surface cooling. Nature, (published online 28 August 2013). Krellenberg, K., A. Mueller, A. Schwarz, R. Hoefer, and J. Welz, 2013: Flood and heat hazards in the Metropolitan Region of Santiago de Chile and the socio-economics of exposure. Applied Geography, 38, 86-95. Krepper, C.M., N.O. García, and P.D. Jones, 2008: Low-frequency response of the upper Paraná basin. International Journal of Climatology, 28(3), 351-360. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Krepper, C.M. and G.V. Zucarelli, 2010a: Climatology of water excesses and shortages in the La Plata Basin. Theoretical and Applied Climatology, 102(1-2), 13-27. Krepper, C.M. and G.V. Zucarelli, 2010b: Climatology of water excesses and shortages in the La Plata Basin. Theoretical and Applied Climatology, 102(1-2), 13-27. Krol, M., A. Jaeger, A. Bronstert, and A. Guentner, 2006: Integrated modelling of climate, water, soil, agricultural and socio-economic processes: A general introduction of the methodology and some exemplary results from the semi-arid north-east of Brazil. Journal of Hydrology, 328(3-4), 417-431. Krol, M.S. and A. Bronstert, 2007: Regional integrated modelling of climate change impacts on natural resources and resource usage in semi-arid Northeast Brazil. Environmental Modelling & Software, 22(2), 259-268. Krol, M.S., M.J. Vries, P.R. Oel, and J.C. Araújo, 2011: Sustainability of Small Reservoirs and Large Scale Water Availability Under Current Conditions and Climate Change. Water Resources Management, 25(12), 3017-3026. Kumar, A., T. Schei, A. Ahenkorah, R.C. Rodriguez, J. Devernay, M. Freitas, D. Hall, A. Killingtveit, and Z. Liu, 2011: Hydropower. In: IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. [Edenhofer, O., R. Pichs-Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 437-496. Kumler, L.M. and M.C. Lemos, 2008: Managing waters of the Paraíba do Sul river basin, Brazil: a case study in institutional change and social learning. Ecology and Society, 13(2), 22. Kundzewicz, Z.W. and P. Döll, 2009: Will groundwater ease freshwater stress under climate change? Hydrological Sciences Journal, 54(4), 665-675. La Rovere, E.L., A.C. Avzaradel, and J.M.G. Monteiro, 2009a: Potential synergy between adaptation and mitigation strategies: production of vegetable oils and biodiesel in northeastern Brazil. Climate Research, 40(2-3), 233- 239. La Rovere, E.L., A.C. Avzaradel, and J.M. Guimaraes Monteiro, 2009b: Potential synergy between adaptation and mitigation strategies: production of vegetable oils and biodiesel in northeastern Brazil. Climate Research, 40(2- 3), 233-239. Lacambra, C. and K. Zahedi, 2011: Climate Change, Natural Hazards and Coastal Ecosystems in Latin-America: A Framework for Analysis. In: Coping with Global Environmental Change, Disasters and Security. Hexagon Series on Human and Environmental Security and Peace Volume 5. [Brauch, H.G., A. Oswald Spring, C. Mesjasz, J. Grin, P. Kameri-Mbote, B. Chourou et al. (eds.)]. Springer Berlin Heidelberg, Berlin, Heidelberg, Germany, pp. 585-601. Lampis, A., 2010: Challenges to Adaptation for Risk-Prone Coastal Livelihoods in Tumaco, Pacific Coast (Colombia). In: Integrative Perspectives on Urbanization and Climate Change. UGEC Viewpoints No. 3. Urbanization and Global Environmental Change (UGEC), pp. 18-22. Langerwisch, F., S. Rost, D. Gerten, B. Poulter, A. Rammig, and W. Cramer, 2013: Potential effects of climate change on inundation patterns in the Amazon Basin. Hydrology and Earth System Sciences, 17(6), 2247-2262. Lapola, D.M., R. Schaldach, J. Alcamo, A. Bondeau, S. Msangi, J.A. Priess, R. Silvestrini, and B.S. Soares, 2011: Impacts of Climate Change and the End of Deforestation on Land Use in the Brazilian Legal Amazon. Earth Interactions, 15(16), 1-29. Lapola, D.M., R. Schaldach, J. Alcamo, A. Bondeau, J. Koch, C. Koelking, and J.A. Priess, 2010: Indirect land-use changes can overcome carbon savings from biofuels in Brazil. Proceedings of the National Academy of Sciences of the United States of America, 107(8), 3388-3393. Lara, A., R. Villalba, and R. Urrutia, 2007: A 400-year tree-ring record of the Puelo River summer fall streamflow in the Valdivian Rainforest eco-region, Chile. Climatic Change, 86(3-4), 331-356. Lardeux, F., P. Loayza, B. Bouchite , and T. Chavez, 2007: Host choice and human blood index of Anopheles pseudopunctipennis in a village of the Andean valleys of Bolivia. Malaria Journal, 6, 8. Larson, A.M., 2010: Making the 'rules of the game': Constituting territory and authority in Nicaragua's indigenous communities. Land use Policy, 27(4), 1143-1152. Laurance, W.F., D.C. Useche, L.P. Shoo, S.K. Herzog, M. Kessler, F. Escobar, G. Brehm, J.C. Axmacher, I.-. Chen, L. Arellano Gamez, P. Hietz, K. Fiedler, T. Pyrcz, J. Wolf, C.L. Merkord, C. Cardelus, A.R. Marshall, C. Ah- Peng, G.H. Aplet, M. del Coro Arizmendi, W.J. Baker, J. Barone, C.A. Bruehl, R.W. Bussmann, D. Cicuzza, G. Eilu, M.E. Favila, A. Hemp, C. Hemp, J. Homeier, J. Hurtado, J. Jankowski, G. Kattan, J. Kluge, T. Kroemer, D.C. Lees, M. Lehnert, J.T. Longino, J. Lovett, P.H. Martin, B.D. Patterson, R.G. Pearson, K.S.-. Peh, B. Richardson, M. Richardson, M.J. Samways, F. Senbeta, T.B. Smith, T.M.A. Utteridge, J.E. Watkins, R. Wilson, Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 S.E. Williams, and C.D. Thomas, 2011: Global warming, elevational ranges and the vulnerability of tropical biota. Biological Conservation, 144(1), 548-557. Lavado, C.W.S., D. Labat, J.L. Guyot, and S. Ardoin-Bardin, 2011: Assessment of climate change impacts on the hydrology of the Peruvian Amazon Andes basin. Hydrological Processes, 25(24), 3721-3734. Lavado, C.W.S., J. Ronchail, D. Labat, J.C. Espinoza, and J.L. Guyot, 2012: Basin-scale analysis of rainfall and runoff in Peru (1969-2004): Pacific, Titicaca and Amazonas drainages. Hydrological Sciences Journal-Journal Des Sciences Hydrologiques, 57(4), 625-642. Lawler, J.J., S.L. Shafer, D. White, P. Kareiva, E.P. Maurer, A.R. Blaustein, and P.J. Bartlein, 2009: Projected climate-induced faunal change in the Western Hemisphere. Ecology, 90(3), 588-597. Le Quesne, C., C. Acuna, J.A. Boninsegna, A. Rivera, and J. Barichivich, 2009: Long-term glacier variations in the Central Andes of Argentina and Chile, inferred from historical records and tree-ring reconstructed precipitation. Palaeogeography Palaeoclimatology Palaeoecology, 281(3-4), 334-344. Leguía, E.J., B. Locatelli, P. Imbach, C.J. Pérez, and R. Vignola, 2008: Servicios Ecosistémicos e Hidroenergía en Costa Rica. Ecosistemas, 17(1), 16-23. Leiva, J.C., G.A. Cabrera, and L.E. Lenzano, 2007: 20 years of mass balances on the Piloto glacier, Las Cuevas river basin, Mendoza, Argentina. Global and Planetary Change, 59(1-4), 10-16. Lemos, M.C., A.R. Bell, N.L. Engle, R.M. Formiga-Johnsson, and D.R. Nelson, 2010: Technical knowledge and water resources management: a comparative study of river basin councils, Brazil. Water Resources Research, 46(6), W06523. Lentes, P., M. Peters, and F. Holmann, 2010: Regionalization of climatic factors and income indicators for milk production in Honduras. Ecological Economics, 69(3), 539-552. Lenton, T.M., H. Held, E. Kriegler, J.W. Hall, W. Lucht, S. Rahmstorf, and H.J. Schellnhuber, 2008: Tipping elements in the Earth's climate system. Proceedings of the National Academy of Sciences of the United States of America, 105(6), 1786-1793. Lesnikowski, A.C., J.D. Ford, L. Berrang-Ford, J.A. Paterson, M. Barrera, and S.J. Heymann, 2011: Adapting to health impacts of climate change: A study of UNFCCC Annex i parties. Environmental Research Letters, 6(4). Lewis, S.L., P.M. Brando, O.L. Phillips, G.M.F. van der Heijden, and D. Nepstad, 2011: The 2010 Amazon Drought. Science, 331(6017), 554-554. Lin, B.B., 2011: Resilience in Agriculture through Crop Diversification: Adaptive Management for Environmental Change. Bioscience, 61(3), 183-193. Loarie, S., D.B. Lobell, G. Asner, Q. Mu, and C. Field, 2011: Direct impacts on local climate of sugar-cane expansion in Brazil. Nature Climate Change, 1, 105-109. Lobell, D.B., W. Schlenker, and J. Costa-Roberts, 2011: Climate trends and global crop production since 1980. Science, 333(6042), 616-620. Lobell, D.B. and C.B. Field, 2007: Global scale climate - crop yield relationships and the impacts of recent warming. Environmental Research Letters, 2(1), 014002. Lobell, D.B., M.B. Burke, C. Tebaldi, M.D. Mastrandrea, W.P. Falcon, and R.L. Naylor, 2008: Prioritizing climate change adaptation needs for food security in 2030. Science, 319(5863), 607-610. Locatelli, B., V. Evans, A. Wardell, A. Andrade, and R. Vignola, 2011: Forests and Climate Change in Latin America: Linking Adaptation and Mitigation. Forests, 2(1), 431-450. Lopes, A.A., A.P. Bandeira, P.C. Flores, and M.V.T. Santana, 2010: Pulmonary hypertension in Latin America: Pulmonary vascular disease: The global perspective. Chest, 137(6 SUPPL.), 78-84. Lopes, C.A., G.O. da Silva, E.M. Cruz, E.D. Assad, and A.d.S. Pereira, 2011: An analysis of the potato production in Brazil upon global warming. Horticultura Brasileira, 29(1), 7-15. Lopez, P., P. Chevallier, V. Favier, B. Pouyaud, F. Ordenes, and J. Oerlemans, 2010: A regional view of fluctuations in glacier length in southern South America. Global and Planetary Change, 71(1-2), 85-108. López, R. and G.I. Galinato, 2007: Should governments stop subsidies to private goods? Evidence from rural Latin America. Journal of Public Economics, 91(5 6), 1071-1094. López-Marrero, T., 2010: An integrative approach to study and promote natural hazards adaptive capacity: A case study of two flood-prone communities in Puerto Rico. Geographical Journal, 176(2), 150-163. Lopez-Rodriguez, S.R. and J.F. Blanco-Libreros, 2008: Illicit crops tropical America: Deforestation, landslides, and the terrestrial carbon stocks. Ambio, 37(2), 141-143. Lorz, C., G. Abbt-Braun, F. Bakker, P. Borges, H. Boernick, L. Fortes, F.H. Frimmel, A. Gaffron, N. Hebben, R. Hoefer, F. Makeschin, K. Neder, L.H. Roig, B. Steiniger, M. Strauch, D. Walde, H. Weiss, E. Worch, and J. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Wummel, 2012: Challenges of an integrated water resource management for the Distrito Federal, Western Central Brazil: climate, land-use and water resources. Environmental Earth Sciences, 65(5), 1575-1586. Losada, I.J., B.G. Reguero, F.J. Méndez, S. Castanedo, A.J. Abascal, and R. Mínguez, 2013: Long-term changes in sea-level components in Latin America and the Caribbean. Global and Planetary Change, 104(0), 34-50. Lowe, R., T.C. Bailey, D.B. Stephenson, R.J. Graham, C.A.S. Coelho, M. Sá Carvalho, and C. Barcellos, 2011: Spatio-temporal modelling of climate-sensitive disease risk: Towards an early warning system for dengue in Brazil. Computers and Geosciences, 37(3), 371-381. Luber, G. and N. Prudent, 2009: Climate change and human health. Transactions of the American Clinical and Climatological Association, 120, 113-117. Lucena, A.F.P., A.S. Szklo, R. Schaeffer, R.R. Souza, B.S. Moreira Cesar Borba, I.V.L. Costa, A.O. Pereira Junior, and S.H.F. Cunha, 2009: The vulnerability of renewable energy to climate change in Brazil. Energy Policy, 37(3), 879-889. Lucena, A.F.P., R. Schaeffer, and A.S. Szklo, 2010a: Least-cost adaptation options for global climate change impacts on the Brazilian electric power system. Global Environmental Change-Human and Policy Dimensions, 20(2), 342-350. Lucena, A.F.P., A.S. Szklo, R. Schaeffer, and R.M. Dutra, 2010b: The vulnerability of wind power to climate change in Brazil. Renewable Energy, 35(5), 904-912. Luque, A., E. Gareth, and C. Lalande, 2013: Climate change governance at the local level: new tools to respond to old deficiencies in Esmeraldas, Ecuador. Local Environment: The International Journal of Justice and Sustainability, (forthcoming). Luzar, J.B., K.M. Silvius, H. Overman, S.T. Giery, J.M. Read, and J.M.V. Fragoso, 2011: Large-scale Environmental Monitoring by Indigenous Peoples. Bioscience, 61(10), 771-781. Lynch, B.D., 2012: Vulnerabilities, competition and rights in a context of climate change toward equitable water governance in Peru's Rio Santa Valley. Global Environmental Change-Human and Policy Dimensions, 22(2), 364-373. Macedo, I.C., J.E.A. Seabra, and J.E.A.R. Silva, 2008: Green house gases emissions in the production and use of ethanol from sugarcane in Brazil: The 2005/2006 averages and a prediction for 2020. Biomass & Bioenergy, 32(7), 582-595. MacNeil, A., S.T. Nichol, and C.F. Spiropoulou, 2011: Hantavirus pulmonary syndrome. Virus Research, 162(1-2), 138-147. Magnan, A., 2009: Proposition d une trame de recherche pour appréhender la capacité d adaptation au changement climatique. Available at: http://vertigo.revues.org/9189. VertigO - La Revue Électronique En Sciences De l'Environnement, 9(3). Magrin, G., C.G. García, D.C. Choque, J.C. Giménez, A.R. Moreno, G.J. Nagy, C. Nobre, and A. Villamizar, 2007a: Latin America. Climate Change 2007: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. In: [Parry, M.L., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, UK, pp. 581-615. Magrin, G.O., M.I. Travasso, W.E. Baethgen, M.O. Grondona, A. Giménez, G. Cunha, J.P. Castano, and G.R. Rodriguez, 2007b: Past and Future Changes in Climate and their Impacts on Annual Crops Yield in South East South America. Available at: http://www.ipcc.ch/pdf/supportingmaterial/tgica_reg-meet-fiji-2007.pdf. In: Meeting Report IPCC TGICA Expert Meeting Integrating Analysis of Regional Climate Change and Response Options. Intergovernmental Panel on Climate Change (IPCC), Nadi,Fiji, pp. 121-124. Magrin, G.O., M.I. Travasso, G.M. López, G.R. Rodríguez, and A.R. Lloveras, 2007c: Vulnerabilidad de la Producción Agrícola en la Región Pampeana Argentina. In: Componente B3 de la 2da Comunicación Nacional de Cambio Climático. Gobierno Argentina, Secretaría de Ambiente y Desarrollo Sustentable, Buenos Aires, Argentina, pp. 84. Magrin, G.O., M.I. Travasso, G.R. Rodríguez, S. Solman, and M. Núnez, 2009: Climate change and wheat production in Argentina. International Journal of Global Warming, 1(1), 214-226. Malhi, Y., J.T. Roberts, R.A. Betts, T.J. Killeen, W. Li, and C.A. Nobre, 2008: Climate change, deforestation, and the fate of the Amazon. Science, 319(5860), 169-172. Malhi, Y., L.E.O.C. Aragao, D. Galbraith, C. Huntingford, R. Fisher, P. Zelazowski, S. Sitch, C. McSweeney, and P. Meir, 2009: Exploring the likelihood and mechanism of a climate-change-induced dieback of the Amazon Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 rainforest. Proceedings of the National Academy of Sciences of the United States of America (PNAS), 106(49), 20610-20615. Mangal, T.D., S. Paterson, and A. Fenton, 2008: Predicting the impact of long-term temperature changes on the epidemiology and control of Schistosomiasis: A mechanistic model. PLoS ONE, 3(1). Mantilla, G., H. Oliveros, and A.G. Barnston, 2009: The role of ENSO in understanding changes in Colombia's annual malaria burden by region, 1960-2006. Malaria Journal, 8(1). Manuel-Navarrete, D., J.J. Gómez, and G. Gallopín, 2007: Syndromes of sustainability of development for assessing the vulnerability of coupled human-environmental systems. The case of hydrometeorological disasters in Central America and the Caribbean. Global Environmental Change-Human and Policy Dimensions, 17(2), 207- 217. Manzello, D.P., J.A. Kleypas, D.A. Budd, C.M. Eakin, P.W. Glynn, and C. Langdon, 2008: Poorly cemented coral reefs of the eastern tropical Pacific: Possible insights into reef development in a high-CO2 world. Proceedings of the National Academy of Sciences of the United States of America, 105(30), 10450-10455. Marcheggiani, S., C. Puccinelli, S. Ciadamidaro, V.D. Bella, M. Carere, M.F. Blasi, N. Pacini, E. Funari, and L. Mancini, 2010: Risks of water-borne disease outbreaks after extreme events. Toxicological and Environmental Chemistry, 92(3), 593-599. Marengo, J.A., 2004: Interdecadal variability and trends of rainfall across the Amazon basin. Theoretical and Applied Climatology, 78(1-3), 79-96. Marengo, J.A., C.A. Nobre, J. Tomasella, M.D. Oyama, G. Sampaio de Oliveira, R. de Oliveira, H. Camargo, L.M. Alves, and I.F. Brown, 2008: The Drought of Amazonia in 2005. Journal of Climate, 21(3), 495-516. Marengo, J.A., R. Jones, L.M. Alves, and M.C. Valverde, 2009a: Future change of temperature and precipitation extremes in South America as derived from the PRECIS regional climate modeling system. International Journal of Climatology, 29(15), 2241-2255. Marengo, J.A., M. Rusticucci, O. Penalba, and M. Renom, 2009b: An intercomparison of observed and simulated extreme rainfall and temperature events during the last half of the twentieth century: part 2: historical trends. Climatic Change, 98(3-4), 509-529. Marengo, J.A., T. Ambrizzi, R. da Rocha, L. Alves, S. Cuadra, M. Valverde, R. Torres, D. Santos, and S. Ferraz, 2010: Future change of climate in South America in the late twenty-first century: intercomparison of scenarios from three regional climate models. Climate Dynamics, 35(6), 1073-1097. Marengo, J.A., S.C. Chou, G. Kay, L.M. Alves, J.F. Pesquero, W.R. Soares, D.C. Santos, A.A. Lyra, G. Sueiro, R. Betts, D.J. Chagas, J.L. Gomes, J.F. Bustamante, and P. Tavares, 2011a: Development of regional future climate change scenarios in South America using the Eta CPTEC/HadCM3 climate change projections: Climatology and regional analyses for the Amazon, Sao Francisco and the Paraná River Basins. Climate Dynamics, 38(9-12), 1829-1848. Marengo, J.A., J.D. Pabón, A. Díaz, G. Rosas, G. Ávalos, E. Montealegre, M. Villacis, S. Solman, and M. Rojas, 2011b: Climate Change: Evidence and Future Scenarios for the Andean Region, Chapter 7. In: Climate Change and Biodiversity in the Tropical Andes. [Herzog, K., R. Martínez, P.M. Jrgensen, and H. Tiessen (eds.)]. Mac Arthur Foundation, IAI, START, Sao Jose dos Campos, Sao Paulo, Brazil, pp. 110-127. Marengo, J.A., J. Tomasella, L.M. Alves, W.R. Soares, and D.A. Rodriguez, 2011c: The drought of 2010 in the context of historical droughts in the Amazon region. Geophysical Research Letters, 38, L12703. Marengo, J.A., J. Tomasella, W.R. Soares, L.M. Alves, and C.A. Nobre, 2012: Extreme climatic events in the Amazon basin Climatological and Hydrological context of recent floods. Theoretical and Applied Climatology, 107(1-2), 73-85. Marengo, J.A., L.M. Alves, W.R. Soares, D.A. Rodriguez, H. Camargo, M. Paredes, and A. Diaz Pablo, 2013a: Two contrasting seasonal extremes in tropical South America in 2012: Flood in Amazonia and drought in Northeast Brazil. Journal of Climate, (doi:10.1175/JCLI-D-12-00642.1, in press). Marengo, J.A., M. Valverde, and G. Obregon, 2013b: Assessments of observed and projected changes in rainfall extremes in the Metropolitan Area of Sao Paulo (MASP). Climate Research, 57, 61-72. Margulis, S., C.B.S. Dubeux, and J. Marcovitch, 2010: Economia da Mudança Climática no Brasil: Custos e Oportunidades. In: . IBEP Gráfica, Sao Paulo, Brazil, pp. 82. Marin, F.R., G.Q. Pellegrino, E.D. Assad, D.S.P. Nassif, M.S. Viana, F.A. Soares, L.L. Cabral, and D. Guiatto, 2009: Cenários futuros para cana-de-açúcar no Estado de Sao Paulo baseados em projeçoes regionalizadas de Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 mudanças climáticas. Proceedings of XVI Congresso Brasileiro de Agrometeorologia, 22-25 September 2009, Gran Darrell Minas Hotel, Eventos e Convençoes Belo Horizonte, Minas Gerais, Brazil, pp. 5. Marín, V.H., A. Tironi, M. Alejandra Paredes, and M. Contreras, 2013: Modeling suspended solids in a Northern Chilean Patagonia glacier-fed fjord: GLOF scenarios under climate change conditions. Ecological Modelling, 264, 7-16. Marini, M.A., M. Barbet-Massin, L.E. Lopes, and F. Jiguet, 2009: Predicted Climate-Driven Bird Distribution Changes and Forecasted Conservation Conflicts in a Neotropical Savanna. Conservation Biology, 23(6), 1558- 1567. Mark, B.G., J. Bury, J.M. McKenzie, A. French, and M. Baraer, 2010: Climate Change and Tropical Andean Glacier Recession: Evaluating Hydrologic Changes and Livelihood Vulnerability in the Cordillera Blanca, Peru. Annals of the Association of American Geographers, 100(4), 794-805. Marques, W.C., 2012: The temporal variability of the freshwater discharge and water levels at the Patos Lagoon, Brazil. International Journal of Geosciences, 3(4), 758-766. Marshall, A., 2012: Existing agbiotech traits continue global march. 30(3), 207-207. Martiello, M.A. and M.V. Giacchi, 2010: Review Article: High temperatures and health outcomes: A review of the literature. Scandinavian Journal of Public Health, 38(8), 826-837. Martínez-Urtaza, J., B. Huapaya, R.G. Gavilan, V. Blanco-Abad, J. Ansede-Bermejo, C. Cadarso-Suarez, A. Figueiras, and J. Trinanes, 2008: Emergence of asiatic vibrio diseases in South America in phase with El Nino. Epidemiology, 19(6), 829-837. Martins, L.D. and M.D.F. Andrade, 2008: Ozone formation potentials of volatile organic compounds and ozone sensitivity to their emission in the megacity of Sao Paulo, Brazil. Water, Air, and Soil Pollution, 195(1-4), 201- 213. Mas-Coma, S., M.A. Valero, and M.D. Bargues, 2009: Climate change effects on trematodiases, with emphasis on zoonotic fascioliasis and schistosomiasis. Veterinary Parasitology, 163(4), 264-280. Masiokas, M.H., R. Villalba, B.H. Luckman, C. Le Quesne, and J.C. Aravena, 2006: Snowpack variations in the central Andes of Argentina and Chile, 1951-2005: Large-scale atmospheric influences and implications for water resources in the region. Journal of Climate, 19(24), 6334-6352. Masiokas, M.H., R. Villalba, B.H. Luckman, M.E. Lascano, S. Delgado, and P. Stepanek, 2008: 20th-century glacier recession and regional hydroclimatic changes in northwestern Patagonia. Global and Planetary Change, 60(1- 2), 85-100. Masiokas, M.H., A. Rivera, L.E. Espizua, R. Villalba, S. Delgado, and J.C. Aravena, 2009: Glacier fluctuations in extratropical South America during the past 1000years. Palaeogeography, Palaeoclimatology, Palaeoecology, 281(3-4), 242-268. Maurer, E., J. Adam, and A. Wood, 2009: Climate model based consensus on the hydrologic impacts of climate change to the Rio Lempa basin of Central America. Hydrology and Earth System Sciences, 13(2), 183-194. McDowell, J.Z. and J.J. Hess, 2012: Accessing adaptation: Multiple stressors on livelihoods in the Bolivian highlands under a changing climate. Global Environmental Change-Human and Policy Dimensions, 22(2), 342- 352. McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks of climate change and human settlements in low elevation coastal zones. Environment and Urbanization, 19(1), 17-37. McGray, H., A. Hammill, and R. Bradley, 2007: Weathering the Storm Options for Framing Adaptation and Development. In: WRI Report . World Resources Institute (WRI), pp. 57. McLeod, E., R. Salm, A. Green, and J. Almany, 2009: Designing marine protected area networks to address the impacts of climate change. Frontiers in Ecology and the Environment, 7(7), 362-370. McMichael, A.J., R.E. Woodruff, and S. Hales, 2006: Climate change and human health: Present and future risks. Lancet, 367(9513), 859-869. McPhee, J., E. Rubio-Alvarez, R. Meza, A. Ayala, X. Vargas, and S. Vicuna, 2010: An Approach to Estimating Hydropower Impacts of Climate Change from a Regional Perspective. [Potter, K.W. and Frevert, D.K. (eds.)]. American Society of Civil Engineers (ASCE), Madison, Wisconsin, USA, pp. 2-2. McSweeney, K. and O.T. Coomes, 2011: Climate-related disaster opens a window of opportunity for rural poor in northeastern Honduras. Proceedings of the National Academy of Sciences of the United States of America, 108(13), 5203-5208. Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Medema, W., B.S. McIntosh, and P.J. Jeffrey, 2008: From premise to practice: a critical assessment of integrated water resources management and adaptive management approaches in the water sector. Ecology and Society, 13(2), 29. Melo, O., X. Vargas, S. Vicuna, F. Meza, and J. McPhee, 2010: Climate Change Economic Impacts on Supply of Water for the M & I Sector in the Metropolitan Region of Chile. [Potter, K.W. and Frevert, D.K. (eds.)]. American Society of Civil Engineers (ASCE), Madison, Wisconsin, USA, pp. 15-15. Mena, N., A. Troyo, R. Bonilla-Carrión, and Ó. Calderón-Arguedas, 2011: Factors associated with incidence of dengue in Costa Rica. Revista Panamericana De Salud Publica/Pan American Journal of Public Health, 29(4), 234-242. Mendes, D. and J.A. Marengo, 2010: Temporal downscaling: a comparison between artificial neural network and autocorrelation techniques over the Amazon Basin in present and future climate change scenarios. Theoretical and Applied Climatology, 100(3-4), 413-421. Menendez, C.G. and A.F. Carril, 2010: Potential changes in extremes and links with the Southern Annular Mode as simulated by a multi-model ensemble. Climatic Change, 98(3-4), 359-377. Mercer, K.L., H.R. Perales, and J.D. Wainwright, 2012: Climate change and the transgenic adaptation strategy: Smallholder livelihoods, climate justice, and maize landraces in Mexico. Global Environmental Change, 22(2), 495-504. Meza, F.J. and D. Silva, 2009: Dynamic adaptation of maize and wheat production to climate change. Climatic Change, 94(1-2), 143-156. Meza, F.J., D. Silva, and H. Vigil, 2008: Climate change impacts on irrigated maize in Mediterranean climates: Evaluation of double cropping as an emerging adaptation alternative. Agricultural Systems, 98(1), 21-30. Meza, F.J., D.S. Wilks, L. Gurovich, and N. Bambach, 2012: Impacts of Climate Change on Irrigated Agriculture in the Maipo Basin, Chile: Reliability of Water Rights and Changes in the Demand for Irrigation. Journal of Water Resources Planning and Management-Asce, 138(5), 421-430. Milne, S. and E. Niesten, 2009: Direct payments for biodiversity conservation in developing countries: practical insights for design and implementation. Oryx, 43(4), 530-541. Minvielle, M. and R.D. Garreaud, 2011: Projecting Rainfall Changes over the South American Altiplano. Journal of Climate, 24(17), 4577-4583. Miteva, D.A., S.K. Pattanayak, and P.J. Ferraro, 2012: Evaluation of biodiversity policy instruments: what works and what doesn t? Oxford Review of Economic Policy, 28(1), 69-92. Mitra, A.K. and G. Rodriguez-Fernandez, 2010: Latin America and the Caribbean: Assessment of the advances in public health for the achievement of the millennium development goals. International Journal of Environmental Research and Public Health, 7(5), 2238-2255. Mittermeier, R.A., P. Robles Gil, and C.G. Mittermeier (eds.), 1997: Megadiversity: Earth s Biologically Wealthiest Nations CEMEX, Monterrey, Mexico. Mittermeier, R.A., P.R. Gil, M. Hoffmann, J. Pilgrim, T. Brooks, C.G. Mittermeier, J. Lamoreux, and G.A.B. Fonseca, 2005: Hotspots revisited: earth's biologically richest and most endangered terrestrial ecoregions. CEMEX, Mexico City, Mexico, 2nd ed., pp. 392. Moncayo, Á. and A.C. Silveira, 2009: Current epidemiological trends for Chagas disease in Latin America and future challenges in epidemiology, surveillance and health policy. Memorias do Instituto Oswaldo Cruz, 104(SUPPL. 1), 17-30. Montagnini, F. and C. Finney, 2011: Payments for Environmental Services in Latin America as a Tool for Restoration and Rural Development. Ambio, 40(3), 285-297. Montenegro, A. and R. Ragab, 2010: Hydrological response of a Brazilian semi-arid catchment to different land use and climate change scenarios: a modelling study. Hydrological Processes, 24(19), 2705-2723. Montenegro, S. and R. Ragab, 2012: Impact of possible climate and land use changes in the semi arid regions: A case study from North Eastern Brazil. Journal of Hydrology, 434-435, 55-68. Monzon, J.P., V.O. Sadras, P.A. Abbate, and O.P. Caviglia, 2007: Modelling management strategies for wheat- soybean double crops in the south-eastern Pampas. Field Crops Research, 101(1), 44-52. Moore, N., E. Arima, R. Walker, and R. Ramos da Silva, 2007: Uncertainty and the changing hydroclimatology of the Amazon. Geophysical Research Letters, 34(14). Mora, C., 2008: A clear human footprint in the coral reefs of the Caribbean. Proceedings of the Royal Society B: Biological Sciences, 275(1636), 767-773. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Moran, E.F., R. Adams, B. Bakoyéma, S.T. Fiorni, and B. Boucek, 2006: Human Strategies for Coping with El Nino Related Drought in Amazônia. Climatic Change, 77(3-4), 343-361. Moreno, A.R., 2006: Climate change and human health in Latin America: Drivers, effects, and policies. Regional Environmental Change, 6(3), 157-164. Moreno, J.E., Y. Rubio-Palis, E. Páez, E. Pérez, and V. Sánchez, 2007: Abundance, biting behaviour and parous rate of anopheline mosquito species in relation to malaria incidence in gold-mining areas of southern Venezuela. Medical and Veterinary Entomology, 21(4), 339-349. Morris, J.N., A.J. Poole, and A.G. Klein, 2006: Retreat of Tropical Glaciers in Colombia and Venezuela from 1984 to 2004 as Measured from ASTER and Landsat Images. Proceedings of 63rd EASTERN SNOW CONFERENCE, 7-9 June 2006, Newark, Delaware, USA. Mosquera-Machado, S. and S. Ahmad, 2006: Flood hazard assessment of Atrato River in Colombia. Water Resources Management, 21(3), 591-609. Moura, R.L.d., C.V. Minte-Vera, I.B. Curado, R.B. Francini Filho, H.d.C.L. Rodrigues, G.F. Dutra, D.C. Alves, and F.J.B. Souto, 2009: Challenges and prospects of fisheries co-management under a marine extractive reserve framework in northeastern Brazil. Coastal Management, 37(6), 617-632. Mueller, A., J. Schmidhuber, J. Hoogeveen, and P. Steduto, 2008: Some insights in the effect of growing bio-energy demand on global food security and natural resources. Water Policy, 10, 83-94. Muggeo, V.M. and S. Hajat, 2009: Modelling the non-linear multiple-lag effects of ambient temperature on mortality in Santiago and Palermo: A constrained segmented distributed lag approach. Occupational and Environmental Medicine, 66(9), 584-591. Mulligan, M., J. Rubiano, G. Hyman, D. White, J. Garcia, M. Saravia, J. Gabriel Leon, J.J. Selvaraj, T. Guttierez, and L. Leonardo Saenz-Cruz, 2010: The Andes basins: biophysical and developmental diversity in a climate of change. Water International, 35(5), 472-492. Murugaiah, C., 2011: The burden of cholera. Critical Reviews in Microbiology, 37(4), 337-348. Nabel, P.E., M. Caretti, and R. Becerra Serial, 2008: Incidencia de aspectos naturales y antrópicos en los anegamientos de la ciudad de Buenos Aires [Incidence of Natural and Anthropic Aspects in the Floods of Buenos Aires City.]. Revista Del Museo Argentino De Ciencias Naturales, 10(1), 37-53. Nabout, J.C., G. Oliveira, M.R. Magalhaes, L.C. Terribile, and F.A. Severo de Almeida, 2011: Global Climate Change and the Production of Pequi Fruits (Caryocar brasiliense) in the Brazilian Cerrado. Natureza & Conservaçao, 9(1), 55-60. Nadal, G.H., G. Bravo, L.O. Girardin, and S. Gortari, 2013: Can renewable energy technologies improve the management of stressed water resources threatened by climate change? Argentine drylands case study. Environment Development and Sustainability, 15(4), 1079-1097. Nakaegawa, T., A. Kitoh, and M. Hosaka, 2013a: Discharge of major global rivers in the late 21st century climate projected with the high horizontal resolution MRI-AGCMs. Hydrological Processes, (published online). Nakaegawa, T., A. Kitoh, H. Murakami, and S. Kusunoki, 2013b: Annual maximum 5-day rainfall total and maximum number of consecutive dry days over Central America and the Caribbean in the late twenty-first century projected by an atmospheric general circulation model with three different horizontal resolutions. Theoretical Applied Climatology, published online, 14. Nakaegawa, T. and W. Vergara, 2010: First Projection of Climatological Mean River Discharges in the Magdalena River Basin, Colombia, in a Changing Climate during the 21st Century. Hydrological Research Letters, 4, 50- 54. Narayan, N., A. Paul, S. Mulitza, and M. Schulz, 2010: Trends in coastal upwelling intensity during the late 20th century. Ocean Science, 6(3), 815-823. Nath, P.K. and B. Behera, 2011: A critical review of impact of and adaptation to climate change in developed and developing economies. Environment, Development and Sustainability, 13(1), 141-162. Nellemann, C., M. MacDevette, T. Manders, B. Eickhout, B. Svihus, A.G. Prins, and B.P. Kaltenborn (eds.), 2009: The Environmental Food Crisis - The Environment's Role in Averting Future Food Crises. Available at: http://www.grida.no/files/publications/FoodCrisis_lores.pdf. In: A UNEP rapid response assessment. United Nations Environment Programme (UNEP), GRID-Arendal, Norway, pp. 104. Nelson, A. and K.M. Chomitz, 2011: Effectiveness of Strict vs. Multiple Use Protected Areas in Reducing Tropical Forest Fires: A Global Analysis Using Matching Methods. Plos One, 6(8), e22722. Nelson, D.R. and T.J. Finan, 2009: Praying for Drought: Persistent Vulnerability and the Politics of Patronage in Ceará, Northeast Brazil. American Anthropologist, 111(3), 302-316. Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Nepstad, D., B.S. Soares-Filho, F. Merry, A. Lima, P. Moutinho, J. Carter, M. Bowman, A. Cattaneo, H. Rodrigues, S. Schwartzman, D.G. McGrath, C.M. Stickler, R. Lubowski, P. Piris-Cabezas, S. Rivero, A. Alencar, O. Almeida, and O. Stella, 2009: The End of Deforestation in the Brazilian Amazon. Science, 326(5958), 1350- 1351. Nepstad, D.C., C.M. Stickler, and O.T. Almeida, 2006: Globalization of the Amazon soy and beef industries: Opportunities for conservation. Conservation Biology, 20(6), 1595-1603. Nepstad, D.C. and C.M. Stickler, 2008: Managing the Tropical Agriculture Revolution. Journal of Sustainable Forestry, 27(1-2), 43-56. Nicholson, L., J. Marin, D. Lopez, A. Rabatel, F. Bown, and A. Rivera, 2009: Glacier inventory of the upper Huasco valley, Norte Chico, Chile: glacier characteristics, glacier change and comparison with central Chile. Annals of Glaciology, 50(53), 111-118. Nobre, C.A. and L.d.S. Borma, 2009: 'Tipping points' for the Amazon forest. Current Opinion in Environmental Sustainability, 1(1), 28-36. Nobre, C.A., 2011: Global Climate Change Modeling: the Brazilian Model of the Global Climate System (MBSCG). In: Research to Advance the Knowledge on Climate Change. FAPESP Research Program on Global Climate Change (FRPGCC). FAPESP, Sao Paulo, Brazil, pp. 8-9. Nobre, C.A., A.D. Young, P.H. Salvida, J.A. Marengo, A.D. Nobre, A. Ogura, O. Thomaz, G.O. Obregon, G.C.M.d. Silva, M. Valverde, A.C. Silveira, and G.O. Rodrigues, 2011: Vulnerabilidade das Megacidades Brasileiras as Mudanças Climáticas: Regiao Metropolitana de Sao Paulo, Relatório Final. INPE-UNICAMP-USP-IPTE- UNESP, Sao Paulo, Brasil, pp. 178. Nóbrega, M.T., W. Collischonn, C.E.M. Tucci, and A.R. Paz, 2011: Uncertainty in climate change impacts on water resources in the Rio Grande Basin, Brazil. Hydrology and Earth System Sciences, 15(2), 585-595. Nogueira, C., P.A. Buckup, N.A. Menezes, O.T. Oyakawa, T.P. Kasecker, M.B. Ramos Neto, and J.M.C. da Silva, 2010: Restricted-Range Fishes and the Conservation of Brazilian Freshwaters. Plos One, 5(6), e11390. Nohara, D., A. Kitoh, M. Hosaka, and T. Oki, 2006: Impact of Climate Change on River Discharge Projected by Multimodel Ensemble. Journal of Hydrometeorology, 7(5), 1076-1089. Nosetto, M.D., E.G. Jobbágy, T. Tóth, and R.B. Jackson, 2008: Regional patterns and controls of ecosystem salinization with grassland afforestation along a rainfall gradient. Global Biogeochemical Cycles, 22(2). Nunez, M.N., S.A. Solman, and M. Fernanda Cabré, 2009: Regional climate change experiments over southern South America. II: Climate change scenarios in the late twenty-first century. Climate Dynamics, 32(7-8), 1081- 1095. Oestreicher, J.S., K. Benessaiah, M.C. Ruiz-Jaen, S. Sloan, K. Turner, J. Pelletier, B. Guay, K.E. Clark, D.G. Roche, M. Meiners, and C. Potvin, 2009: Avoiding deforestation in Panamanian protected areas: An analysis of protection effectiveness and implications for reducing emissions from deforestation and forest degradation. Global Environmental Change-Human and Policy Dimensions, 19(2), 279-291. Oft, P., 2010: Micro-Finance Instruments Can Contribute to Build Resilience. A Case Study of Coping and Adaptation Strategies to Climate-Related Shocks in Piura, Peru. In: Graduate Research Series. PhD Dissertations. Publication Series of UNU-EHS Vol. 2. UNU-EHS, Bonn, Germany, pp. 157. Oliveira, B.F.A.d., E. Ignotti, and S.S. Hacon, 2011: A systematic review of the physical and chemical characteristics of pollutants from biomass burning and combustion of fossil fuels and health effects in Brazil. Cadernos De Saude Publica, 27(9), 1678-1698. Oliveira, P.J.C., G.P. Asner, D.E. Knapp, A. Almeyda, R. Galvan-Gildemeister, S. Keene, R.F. Raybin, and R.C. Smith, 2007: Land-use allocation protects the Peruvian Amazon. Science, 317(5842), 1233-1236. Olmo, N.R.S., P.H.N. Saldiva, A.L.F. Braga, C.A. Lin, U.P. Santos, and L.A.A. Pereirai, 2011: A review of low- level air pollution and adverse effects on human health: Implications for epidemiological studies and public policy. Clinics, 66(4), 681-690. Olson, S.H., R. Gangnon, E. Elguero, L. Durieux, J.-. Guégan, J.A. Foley, and J.A. Patz, 2009: Links between climate, malaria, and wetlands in the amazon basin. Emerging Infectious Diseases, 15(4), 659-662. Oltremari, J.V. and R.G. Jackson, 2006: Conflicts, perceptions, and expectations of indigenous communities associated with natural areas in Chile. Natural Areas Journal, 26(2), 215-220. Osorio, L., J. Todd, R. Pearce, and D.J. Bradley, 2007: The role of imported cases in the epidemiology of urban Plasmodium falciparum malaria in Quibdó, Colombia. Tropical Medicine and International Health, 12(3), 331- 341. Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ospina-Norena, J.E., C. Gay-García, A.C. Conde, M.A.G. Ana V, and G. Sánchez-Torres Esqueda, 2009a: Vulnerability of water resources in the face of potential climate change: generation of hydroelectric power in Colombia. Atmosfera, 22(3), 229. Ospina-Norena, J.E., C. Gay-García, A.C. Conde, and G. Sánchez-Torres Esqueda, 2009b: Analysis of the water supply-demand relationship in the Sinú-Caribe basin, Colombia, under different climate change scenarios. Atmosfera, 22(4), 399-412. Ospina-Norena, J.E., C. Gay-García, A.C. Conde, and G. Sánchez-Torres Esqueda, 2011a: Water availability as a limiting factor and optimization of hydropower generation as an adaptation strategy to climate change in the Sinú-Caribe river basin. Atmosfera, 24(2), 203. Ospina-Norena, J.E., C. Gay-García, A.C. Conde, and G. Sánchez-Torres Esqueda, 2011b: A proposal for a vulnerability index for hydroelectricity generation in the face of potential climate change in Colombia. Atmosfera, 24(3), 329. Palmer, M.A., C.A. Reidy Liermann, C. Nilsson, M. Flörke, J. Alcamo, P.S. Lake, and N. Bond, 2008: Climate change and the world's river basins: anticipating management options. Frontiers in Ecology and the Environment, 6(2), 81-89. Pasquini, A.I., K.L. Lecomte, E.L. Piovano, and P.J. Depetris, 2006: Recent rainfall and runoff variability in central Argentina. Quaternary International, 158(1), 127-139. Pasquini, A.I. and P.J. Depetris, 2007: Discharge trends and flow dynamics of South American rivers draining the southern Atlantic seaboard: An overview. Journal of Hydrology, 333(2-4), 385-399. Pasquini, A.I., K.L. Lecomte, and P.J. Depetris, 2008: Climate change and recent water level variability in Patagonian proglacial lakes, Argentina. Global and Planetary Change, 63(4), 290-298. Pasquini, A.I., K.L. Lecomte, and P.J. Depetris, 2013: The Manso Glacier drainage system in the northern Patagonian Andes: an overview of its main hydrological characteristics. Hydrological Processes, 27(2), 217- 224. Payne, L. and J.R. Fitchett, 2010: Bringing neglected tropical diseases into the spotlight. Trends in Parasitology, 26(9), 421-423. Pellicciotti, F., P. Burlando, and K. Van Vliet, 2007: Recent trends in precipitation and streamflow in the Aconcagua River basin, central Chile[Ginot P, S.J. (ed.)]. Proceedings of Proceedings of a workshop on Andean Glaciology and a symposium on the Contribution from Glaciers and Snow Cover to Runoff from Mountains in Different Climates during the 7th Scientific Assembly of the International Association of Hydrological Sciences (IAHS), 4-9 April 2005, Foz do Iguacu, Brazil. Penalba, O.C. and J.A. Rivera, 2013: Future changes in drought characteristics over Southern South America projected by a CMIP5 multi-model ensemble. American Journal of Climate Change, 2(3). Penalba, O.C. and F.A. Robledo, 2010: Spatial and temporal variability of the frequency of extreme daily rainfall regime in the La Plata Basin during the 20th century. Climatic Change, 98(3-4), 531-550. Peraza, S., C. Wesseling, A. Aragon, R. Leiva, R.A. García-Trabanino, C. Torres, K. Jakobsson, C.G. Elinder, and C. Hogstedt, 2012: Decreased Kidney Function Among Agricultural Workers in El Salvador. American Journal of Kidney Diseases, 59(4), 531-540. Perazzoli, M., A. Pinheiro, and V. Kaufmann, 2013: Assessing the impact of climate change scenarios on water resources in southern Brazil. Hydrological Sciences Journal, 58(1), 77-87. Pereira, L.F.M., S. Barreto, and J. Pittock, 2009: Participatory river basin management in the Sao Joao River, Brazil: A basis for climate change adaptation? Climate and Development, 1(3), 261-268. Perera, F.P., 2008: Children are likely to suffer most from our fossil fuel addiction. Environmental Health Perspectives, 116(8), 987-990. Pérez, C., C. Nicklin, O. Dangles, S. Vanek, S. Sherwood, S. Halloy, K. Garrett, and G. Forbes, 2010: Climate Change in the High Andes: Implications and Adaptation Strategies for Small-scale Farmers. The International Journal of Environmental, Cultural, Economic and Social Sustainability, 6(5), 71-88. Peterson, M.J., D.M. Hall, A.M. Feldpausch-Parker, and T.R. Peterson, 2010: Obscuring Ecosystem Function with Application of the Ecosystem Services Concept. Conservation Biology, 24(1), 113-119. Pettengell, C., 2010: Adaptación al cambio climático. Capacitar a las personas que viven en la pobreza para que puedan adaptarse. In: Informe de Investigación de OXFAM. OXFAM, UK, pp. 56. Pielke Jr, R.A., J. Rubiera, C. Landsea, M.L. Fernández, and R. Klein, 2003: Hurricane vulnerability in Latin America and the Caribbean: Normalized damage and loss potentials. Natural Hazards Review, 4(3), 101-114. Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pinault, L.L. and F.F. Hunter, 2011: New highland distribution records of multiple Anopheles species in the Ecuadorian Andes. Malaria Journal, 10, 236. Pinto, O.,Jr. and I.R.C.A. Pinto, 2008: On the sensitivity of cloud-to-ground lightning activity to surface air temperature changes at different timescales in Sao Paulo, Brazil. Journal of Geophysical Research Atmospheres, 113, D20123. Pinto, H.S., E.D. Assad, J.Z. Junior, S.R.M. Evangelista, A.F. Otavian, A.M.H. Ávila, B. Evangelista, F.R. Marin, C.M. Junior, G.Q. Pellegrino, P.P. Coltri, and G. Coral, 2008: Aquecimento global e a nova geografia da produçao agrícola no Brasil. Embrapa/Unicamp, Sao Paulo, Brazil, pp. 81. Pinto, O., I.R.C.A. Pinto, and M.A.S. Ferro, 2013: A study of the long-term variability of thunderstorm days in southeast Brazil. Journal of Geophysical Research: Atmospheres, 118(11), 5231-5246. Pittock, J., 2011: National Climate Change Policies and Sustainable Water Management: Conflicts and Synergies. Ecology and Society, 16(2), 25. Pizarro, R., F. Balocchi, M. Vera, A. Aguilera, C. Morales, R. Valdés, C. Sangüesa, C. Vallejos, R. Fuentes, A. Abarza, and C. Olivares, 2013: Influencia del cambio climático en el comportamiento de los caudales máximos en la zona mediterránea de Chile. Tecnología y Ciencias Del Agua, 4(2), 5-19. Plaza, G. and M. Pasculi, 2007: Estrategias de adaptación al cambio climático: caso de estudio de la localidad de Aguaray Salta [Adaptation strategies to climate change: case study of the town of Salta Aguaray]. Avances En Energías Renovables y Medio Ambiente, 11, 129-136. PNCC, 2007: Vulnerabilidad y Adaptación al Cambio Climático en Bolivia. Resultados de un proceso de investigación participativa en las regiones del Lago Titicaca y los Valles Crucenos. UNDP; República de Bolivia, Programa Nacional de Cambios Climáticos (PNCC), pp. 141. Podestá, G., F. Bert, B. Rajagopalan, S. Apipattanavis, C. Laciana, E. Weber, W. Easterling, R. Katz, D. Letson, and A. Menendez, 2009: Decadal climate variability in the Argentine Pampas: regional impacts of plausible climate scenarios on agricultural systems. Climate Research, 40(2-3), 199-210. Polidoro, B.A., K.E. Carpenter, L. Collins, N.C. Duke, A.M. Ellison, J.C. Ellison, E.J. Farnsworth, E.S. Fernando, K. Kathiresan, N.E. Koedam, S.R. Livingstone, T. Miyagi, G.E. Moore, Vien Ngoc Nam, J.E. Ong, J.H. Primavera, S.G. Salmo III, J.C. Sanciangco, S. Sukardjo, Y. Wang, and J.W.H. Yong, 2010: The Loss of Species: Mangrove Extinction Risk and Geographic Areas of Global Concern. Plos One, 5(4), e10095. Polissar, P.J., M.B. Abbott, A.P. Wolfe, M. Bezada, V. Rull, and R.S. Bradley, 2006: Solar modulation of Little Ice Age climate in the tropical Andes. Proc Natl Acad Sci U S A, 103(24), 8937-42. Porter-Bolland, L., E.A. Ellis, M.R. Guariguata, I. Ruiz-Mallén, S. Negrete-Yankelevich, and V. Reyes-García, 2012: Community managed forests and forest protected areas: An assessment of their conservation effectiveness across the tropics. Forest Ecology and Management, 268(0), 6-17. Poveda, G. and K. Pineda, 2009: Reassessment of Colombia s tropical glaciers retreat rates: are they bound to disappear during the 2010 2020 decade? Advances in Geosciences, 22, 107-116. Poveda, G., Ó.A. Estrada-Restrepo, J.E. Morales, Ó.O. Hernández, A. Galeano, and S. Osorio, 2011: Integrating knowledge and management regarding the climate-malaria linkages in Colombia. Current Opinion in Environmental Sustainability, 3(6), 449-460. Programa Estado de la Nación-Región, 2011: Cuarto Informe Estado de la Región en Desarrollo Humano Sostenible., Costa Rica, pp. 550. Przeslawski, R., S. Ahyong, M. Byrne, G. Woerheide, and P. Hutchings, 2008: Beyond corals and fish: the effects of climate change on noncoral benthic invertebrates of tropical reefs. Global Change Biology, 14(12), 2773-2795. Quintana, J.M. and P. Aceituno, 2012: Changes in the rainfall regime along the extratropical west coast of South America (Chile): 30-43 S. Atmósfera, 25(1), 1-12. Quiroga, A. and C. Gaggioli, 2011: Gestión del agua y viabilidad de los sistemas productivos. In: Condiciones para el Desarrollo de Producciones Agrícola-Ganaderas en el SO Bonaerense. Academia Nacional de Agronomía y Veterinaria de la República Argentina, Tomo LXIV, Buenos Aires, Argentina, pp. 233-249. Rabatel, A., H. Castebrunet, V. Favier, L. Nicholson, and C. Kinnard, 2011: Glacier changes in the Pascua-Lama region, Chilean Andes (29° S): recent mass balance and 50 yr surface area variations. The Cryosphere, 5(4), 1029-1041. Rabatel, A., B. Francou, A. Soruco, J. Gomez, B. Cáceres, J.L. Ceballos, R. Basantes, M. Vuille, J.-. Sicart, C. Huggel, M. Scheel, Y. Lejeune, Y. Arnaud, M. Collet, T. Condom, G. Consoli, V. Favier, V. Jomelli, R. Galarraga, P. Ginot, L. Maisincho, J. Mendoza, M. Ménégoz, E. Ramirez, P. Ribstein, W. Suarez, M. Villacis, Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and P. Wagnon, 2013: Current state of glaciers in the tropical Andes: a multi-century perspective on glacier evolution and climate change. The Cryosphere Discussions, 7, 81-102. Rabatel, A., M. A., B. Francou, and V. Jomelli, 2006: Glacier recession on Cerro Charquini (16 degrees S), Bolivia, since the maximum of the Little Ice Age (17th century). Journal of Glaciology, 52(176), 110. Rabatel, A., B. Francou, V. Jomelli, P. Naveau, and D. Grancher, 2008: A chronology of the Little Ice Age in the tropical Andes of Bolivia (16°S) and its implications for climate reconstruction. Quaternary Research, 70(2), 198-212. Racoviteanu, A.E., W.F. Manley, Y. Arnaud, and M.W. Williams, 2007: Evaluating digital elevation models for glaciologic applications: An example from Nevado Coropuna, Peruvian Andes. Global and Planetary Change, 59(1-4), 110-125. Radachowsky, J., V.H. Ramos, R. McNab, E.H. Baur, and N. Kazakov, 2012: Forest concessions in the Maya Biosphere Reserve, Guatemala: A decade later. Forest Ecology and Management, 268, 18-28. Ramankutty, N., H.K. Gibbs, F. Achard, R. Defriess, J.A. Foley, and R.A. Houghton, 2007: Challenges to estimating carbon emissions from tropical deforestation. Global Change Biology, 13(1), 51-66. Ramirez-Villegas, J., M. Salazar, A. Jarvis, and C.E. Navarro-Racines, 2012: A way forward on adaptation to climate change in Colombian agriculture: perspectives towards 2050. Climatic Change, 115(3-4), 611-628. Rammig, A., T. Jupp, K. Thonicke, B. Tietjen, J. Heinke, S. Ostberg, W. Lucht, W. Cramer, and P. Cox, 2010: Estimating the risk of Amazonian forest dieback. New Phytologist, 187(3), 694-706. Raup, B., A. Racoviteanu, S.J.S. Khalsa, C. Helm, R. Armstrong, and Y. Arnaud, 2007: The GLIMS geospatial glacier database: A new tool for studying glacier change. Global and Planetary Change, 56(1-2), 101-110. Ravera, F., D. Tarrason, and E. Simelton, 2011: Envisioning Adaptive Strategies to Change: Participatory Scenarios for Agropastoral Semiarid Systems in Nicaragua. Ecology and Society, 16(1). Re, M. and V.R. Barros, 2009: Extreme rainfalls in SE South America. Climatic Change, 96(1-2), 119-136. Ready, P.D., 2008: Leishmaniasis emergence and climate change. OIE Revue Scientifique Et Technique, 27(2), 399- 412. Rebaudo, F., V. Crespo-Pérez, J. Silvain, and O. Dangles, 2011: Agent-Based Modeling of Human-Induced Spread of Invasive Species in Agricultural Landscapes: Insights from the Potato Moth in Ecuador. Jasss-the Journal of Artificial Societies and Social Simulation, 14(3), 7. República Argentina, 2007: 2da Comunicación Nacional de la República Argentina a la Convención Marco de las Naciones Unidas sobre Cambio Climático. República Argentina;, Buenos Aires, Argentina, pp. 201. Restrepo, C. and N. Alvarez, 2006: Landslides and their contribution to land-cover change in the mountains of Mexico and Central America. Biotropica, 38(4), 446-457. Restrepo-Pineda, E., E. Arango, A. Maestre, V.E.D. Rosário, and P. Cravo, 2008: Studies on antimalarial drug susceptibility in Colombia, in relation to Pfmdr1 and Pfcrt. Parasitology, 135(5), 547-553. Reuveny, R., 2007: Climate change-induced migration and violent conflict. Political Geography, 26(6), 656-673. Rivarola Sosa, J.M., G. Brandani, C. Dibari, M. Moriondo, R. Ferrise, G. Trombi, and M. Bindi, 2011: Climate change impact on the hydrological balance of the Itaipu Basin. Meteorological Applications, 18(2), 163-170. Rivera, J.A., O.C. Penalba, and M.L. Betolli, 2013: Inter-annual and inter-decadal variability of dry days in Argentina. International Journal of Climatology, 33(4), 834-842. Roberts, N., 2009: Culture and landslide risk in the Central Andes of Bolivia and Peru. Studia UBB Geologia, 54(1), 55-59. Rodrigues Capítulo, A., N. Gómez, A. Giorgi, and C. Feijoó, 2010: Global changes in pampean lowland streams (Argentina): implications for biodiversity and functioning. Hydrobiologia, 657(1), 53-70. Rodrigues, R.R., S. Gandolfi, A.G. Nave, J. Aronson, T.E. Barreto, C.Y. Vidal, and P.H.S. Brancalion, 2011: Large- scale ecological restoration of high-diversity tropical forests in SE Brazil. Forest Ecology and Management, 261(10), 1605-1613. Rodríguez Laredo, D.M., 2011: La gestión del verde urbano como un criterio de mitigación y adaptación al cambio climatico Revista De La Asociación Argentina De Ecología De Paisajes, 2(2), 123-130. Rodríguez, J.P., F. Rojas-Suárez, and D. Hernández (eds.), 2010: Libro Rojo de los Ecosistemas Terrestres de Venezuela Provita, Venezuela, pp. 324. Rodriguez, A., M. Vaca, G. Oviedo, S. Erazo, M.E. Chico, C. Teles, M.L. Barreto, L.C. Rodrigues, and P.J. Cooper, 2011: Urbanisation is associated with prevalence of childhood asthma in diverse, small rural communities in Ecuador. Thorax, 66(12), 1043-1050. Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rodriguez, D.A., J. Tomasella, and C. Linhares, 2010: Is the forest conversion to pasture affecting the hydrological response of Amazonian catchments? Signals in the Ji-Paraná Basin. Hydrological Processes, 24(10), 1254- 1269. Rodríguez-Morales, A. and A. Herrera-Martinez, 2009: Potential influence of climate variability on dengue incidence in a western pediatric hospital of Venezuela, 2001-2008. Tropical Medicine & International Health, 14, 164-165. Rodríguez-Morales, A.J., L. Delgado, N. Martinez, and C. Franco-Paredes, 2006: Impact of imported malaria on the burden of disease in northeastern Venezuela. Journal of Travel Medicine, 13(1), 15-20. Rodríguez-Morales, A.J., L. Rada, J. Benitez, and C. Franco-Paredes, 2007: Impact of climate variability on cutaneous leishmaniasis in Venezuela. American Journal of Tropical Medicine and Hygiene, 77(5), 228-229. Rodríguez-Morales, A.J., L. Echezuria, and A. Risquez, 2010: Impact of Climate Change on Health and Disease in Latin America, Climate Change and Variability. In: Climate Change and Variability. [Simard, S. (ed.)]. Sciyo, , pp. 464-486. Rodríguez-Morales, A.J., 2011: Cambio Climático, precipitaciones, sociedad y desastres en América Latina: relaciones y necesidades [Climate change, rainfall, society and disasters in Latin America: relations and needs]. Revista Peruana De Medicina Experimental y Salud Publica, 28(1), 165-6. Rodríguez-Pérez, M.A., T.R. Unnasch, and O. Real-Najarro, 2011: Assessment and Monitoring of Onchocerciasis in Latin America. Advances in Parasitology, 77, 175-226. Roebeling, P.C. and E.M.T. Hendrix, 2010: Land speculation and interest rate subsidies as a cause of deforestation: The role of cattle ranching in Costa Rica. Land use Policy, 27(2), 489-496. Romero, G.A.S. and M. Boelaert, 2010: Control of Visceral Leishmaniasis in Latin America-A Systematic Review. Plos Neglected Tropical Diseases, 4(1), e584. Romero-Lankao, P., 2007a: Are we missing the point? Particularities of urbanization, sustainability and carbon emissions in Latin American cities. Environment and Urbanization, 19(1), 159-175. Romero-Lankao, P., 2007b: How do Local Governments in Mexico City Manage Global Warming? Local Environment, 12(5), 519-535. Romero-Lankao, P., 2010: Water in Mexico City: what will climate change bring to its history of water-related hazards and vulnerabilities? Environment and Urbanization, 22(1), 157-178. Romero-Lankao, P., H. Qin, S. Hughes, M. Haeffner, and M. Borbor-Cordova, 2012: Urban Vulnerability and Adaptation to the Health Impacts of Air Pollution and Climate Extremes in Latin American Cities. Research in Urban Sociology, 12, 247-275. Romero-Lankao, P., S. Hughes, A. Rosas-Huerta, and Borquez,R.,Gnatz,D., 2013a: Urban Institutional Response Capacity for Climate Change: An examination of construction and pathways in Mexico City and Santiago. Environment and Planning C, (forthcoming). Romero-Lankao, P., H. Qin, and M. Borbor-Cordova, 2013b: Exploration of health risks related to air pollution and temperature in three Latin American cities Social Sciences and Medicine, 83, 110-118. Romero-Lankao, P. and H. Qin, 2011: Conceptualizing urban vulnerability to global climate and environmental change. Current Opinion in Environmental Sustainability, 3(3), 142-149. Romero-Lankao, P., 2012: Governing Carbon and Climate in the Cities: An Overview of Policy and Planning Challenges and Options. European Planning Studies, 20(1), 7-26. Romero-Lankao, P., M. Borbor-Cordova, R. Abrutsky, G. Günther, E. Behrenz, and L. Dawidowsky, 2012a: ADAPTE: A tale of diverse teams coming together to do issue-driven interdisciplinary research. Environmental Science & Policy, 26, 29-39. Romero-Lankao, P., H. Qin, and K. Dickinson, 2012b: Urban vulnerability to temperature-related hazards: A meta- analysis and meta-knowledge approach. Global Environmental Change, 22(3), 670-683. Roncoli, C., 2006: Ethnographic and participatory approaches to research on farmers' responses to climate predictions. Climate Research, 33(1), 81-99. Rotureau, B., P. Couppié, M. Nacher, J.-. Dedet, and B. Carme, 2007: Cutaneous leishmaniases in French Guiana. Bulletin De La Societe De Pathologie Exotique, 100(4), 251-260. Ruane, A.C., L.D. Cecil, R.M. Horton, R. Gordón, R. McCollum, D. Brown, B. Killough, R. Goldberg, A.P. Greeley, and C. Rosenzweig, 2013: Climate change impact uncertainties for maize in Panama: Farm information, climate projections, and yield sensitivities. Agricultural and Forest Meteorology, 170(0), 132-145. Rubin, O. and T. Rossing, 2012: National and Local Vulnerability to Climate-Related Disasters in Latin America: The Role of Social Asset-Based Adaptation. Bulletin of Latin American Research, 31(1), 19-35. Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rubio-Álvarez, E. and J. McPhee, 2010: Patterns of spatial and temporal variability in streamflow records in south central Chile in the period 1952 2003. Water Resources Research, 46(5), W05514. Ruiz, D., H.A. Moreno, M.E. Gutiérrez, and P.A. Zapata, 2008: Changing climate and endangered high mountain ecosystems in Colombia. Science of the Total Environment, 398(1-3), 122-132. Rusticucci, M. and M. Renom, 2008: Variability and trends in indices of quality-controlled daily temperature extremes in Uruguay. International Journal of Climatology, 28(8), 1083-1095. Rusticucci, M. and B. Tencer, 2008: Observed Changes in Return Values of Annual Temperature Extremes over Argentina. Journal of Climate, 21(21), 5455-5467. Rusticucci, M., 2012: Observed and simulated variability of extreme temperature events over South America. Atmospheric Research, 106, 1-17. Salazar, L.F., C.A. Nobre, and M.D. Oyama, 2007: Climate change consequences on the biome distribution in tropical South America. Geophysical Research Letters, 34(9), L09708. Salazar-Lindo, E., C. Seas, and D. Gutierrez, 2008: ENSO and cholera in South America: What can we learn about it from the 1991 cholera outbreak? International Journal of Environment and Health, 2(1), 30-36. Saldana-Zorrilla, S.O., 2008: Stakeholders views in reducing rural vulnerability to natural disasters in Southern Mexico: Hazard exposure and coping and adaptive capacity. Global Environmental Change, 18(4), 583-597. Salinas, H., J. Almenara, Á. Reyes, P. Silva, M. Erazo, and M.J. Abellán, 2006: Estudio de variables asociadas al cáncer de piel en Chile mediante análisis de componentes principales [Study of variables associated with skin cancer in Chile using principal component analysis]. Actas Dermo-Sifiliográficas, 97(4), 241-246. Salomón, O.D., Y. Basmajdian, M.S. Fernández, and M.S. Santini, 2011: Lutzomyia longipalpis in Uruguay: The first report and the potential of visceral leishmaniasis transmission. Memorias do Instituto Oswaldo Cruz, 106(3), 381-382. Salzmann, N., C. Huggel, P. Calanca, A. Díaz, T. Jonas, C. Jurt, T. Konzelmann, P. Lagos, M. Rohrer, W. Silverio, and M. Zappa, 2009: Integrated assessment and adaptation to climate change impacts in the Peruvian Andes. Advances in Geosciences, 22, 35-39. Salzmann, N., C. Huggel, M. Rohrer, W. Silverio, B.G. Mark, P. Burns, and C. Portocarrero, 2013: Glacier changes and climate trends derived from multiple sources in the data scarce Cordillera Vilcanota region, southern Peruvian Andes. Cryosphere, 7(1), 103-118. Sampaio, G., C. Nobre, M.H. Costa, P. Satyamurty, B.S. Soares-Filho, and M. Cardoso, 2007: Regional climate change over eastern Amazonia caused by pasture and soybean cropland expansion. Geophysical Research Letters, 34(17), L17709. Sankarasubramanian, A., U. Lall, F.A. Souza Filho, and A. Sharma, 2009: Improved water allocation utilizing probabilistic climate forecasts: Short-term water contracts in a risk management framework. Water Resources Research, 45, W11409. Sansigolo, C.A. and M.T. Kayano, 2010: Trends of seasonal maximum and minimum temperatures and precipitation in Southern Brazil for the 1913-2006 period. Theoretical and Applied Climatology, 101(1-2), 209-216. Santos, C.A. and J.I.B. Brito, 2007: Análise dos índices de extremos para o semiárido do Brasil e suas relaçoes com TSM e IVDN [Analysis of indices of extremes for the semi-arid region of Brazil and its relationship with SST and NDVI]. Revista Brasileira De Meteorologia, 22(3), 303-312. Santos, C.A.C., J.I.B. Brito, T.V.R. Rao, and E.A. Meneses, 2009: Tendencias dos Índices de precipitaçao no Estado do Ceará. Revista Brasileira De Meteorologia, 24(1), 39-47. Sathaye, J., O. Lucon, A. Rahman, J. Christensen, F. Denton, J. Fujino, G. Heath, M. Mirza, H. Rudnick, A. Schlaepfer, and A. Shmakin, 2011: Renewable Energy in the Context of Sustainable Development. In: IPCC Special Report on Renewable Energy Sources and Climate change Mitigation. [Edenhofer, O., R. Pichs- Madruga, Y. Sokona, K. Seyboth, P. Matschoss, S. Kadner et al. (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 707-790. Satterthwaite, D., 2011a: How urban societies can adapt to resource shortage and climate change. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1942), 1762-1783. Satterthwaite, D., 2011b: How can urban centers adapt to climate change with ineffective or unrepresentative local governments? Wiley Interdisciplinary Reviews-Climate Change, 2(5), 767-776. Satyamurty, P., C.P.W. da Costa, A.O. Manzi, and L.A. Candido, 2013: A quick look at the 2012 record flood in the Amazon Basin. Geophysical Research Letters, 40(7), 1396-1401. Satyamurty, P., A.A. de Castro, J. Tota, L.E. da Silva Gularte, and A.O. Manzi, 2010: Rainfall trends in the Brazilian Amazon Basin in the past eight decades. Theoretical and Applied Climatology, 99(1-2), 139-148. Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Saulo, C., L. Ferreira, J. Nogués-Paegle, M. Seluchi, and J. Ruiz, 2010: Land-Atmosphere Interactions during a Northwestern Argentina Low Event. Monthly Weather Review, 138(7), 2481-2498. Saurral, R.I., V.R. Barros, and D.P. Lettenmaier, 2008: Land use impact on the Uruguay River discharge. Geophysical Research Letters, 35(12), L12401. Sayago, J.M., M.M. Collantes, L.d.V. Neder, and J. Busnelli, 2010: Cambio climático y amenazas ambientales en el Área Metropolitana de Tucumán [ Climate change and environmental hazard at the Metropolitan Area of Tucumán]. Revista De La Asociación Geológica Argentina, 66(4), 544-554. Schneider, C., M. Schnirch, C. Acuna, G. Casassa, and R. Kilian, 2007: Glacier inventory of the Gran Campo Nevado Ice Cap in the Southern Andes and glacier changes observed during recent decades. Global and Planetary Change, 59(1-4), 87-100. Schulz, N., J.P. Boisier, and P. Aceituno, 2012: Climate change along the arid coast of northern Chile. International Journal of Climatology, 32(12), 1803-1814. Scott, C.A., R.G. Varady, F. Meza, E. Montana, G.B. de Raga, B. Luckman, and C. Martius, 2012: Science-Policy Dialogues for Water Security: Addressing Vulnerability and Adaptation to Global Change in the Arid Americas. Environment, 54(3), 30-42. Scott, C.A., F.J. Meza, R.G. Varady, H. Tiessen, J. McEvoy, G.M. Garfin, M. Wilder, L.M. Farfan, N. Pineda Pablos, and E. Montana, 2013: Water Security and Adaptive Management in the Arid Americas. Annals of the Association of American Geographers, 103(2), 280-289. Seiler, C., R.W.A. Hutjes, and P. Kabat, 2013: Likely Ranges of Climate Change in Bolivia. Journal of Applied Meteorology and Climatology, 52(6), 1303-1317. SENAMHI, 2005: Escenarios del cambio climático en el Perú al 2050 - cuenca del río Piura. [Climate change scenarios in Peru for 2050 the Piura River basin.]. Programa de Cambio Climático y Calidad de Aire, Servicio Nacional de Meteorología e Hidrología, Lima, Perú. Segunda edición: octubre 2005, pp. 197. SENAMHI, 2007: Escenarios de cambio climático en la Cuenca del Río Urubamba para el ano 2100 [Climate change scenarios in the Urubamba River Basin by 2100]. [Rosas, G., Avalos, G., Díaz, A., Oria, C., Acuna, D., Metzger, L. and Miguel, R. (eds.)]. Servicio Nacional de Meteorología e Hidrología (SENAMHI), Lima, Perú. Segunda edición: octubre de 2005, pp. 120. SENAMHI, 2009a: Escenarios climáticos en la cuenca del rio Mayo para el ano 2030 [Climate scenarios in the Mayo River Basin by the year 2030] Available at: http://www.senamhi.gob.pe. [Obregón, G., Diaz, A., Rosas, G., Avalos, G., Oria, C., Acuna, D., Llacza, A. and Miguel, R. (eds.)]. Servicio Nacional de Meteorología e Hidrología (SENAMHI); Centro de Predicción Numérica (CPN), pp. 133. SENAMHI, 2009b: Escenarios climáticos en la cuenca del rio Santa para el ano 2030 [Climate scenarios in the Santa River Basin by 2030]. Available at: http://www.senamhi.gob.pe. In: SENAMHI (2009d) Escenarios Climáticos em la Cuenca del Rio Santa para 2030. SENAMHI Servicio Nacional de Meteorología e Hidrología- Centro de Predicción Numérica CP, , 139 pp. [Obregón, G., Diaz, A., Rosas, G., Avalos, G., Oria, C., Acuna, D., Llacza, A. and Miguel, R. (eds.)]. Servicio Nacional de Meteorología e Hidrología (SENAMHI); Centro de Predicción Numérica (CPN), pp. 139. SENAMHI, 2009c: Escenarios de Cambio Climático en la Cuenca del Río Mantaro para 2100 [Climate Change Scenarios in the Mantaro River Basin by 2100]. Available at: http://www.senamhi.gob.pe. Servicio Nacional de Meteorología e Hidrología (SENAMHI); Centro de Predicción Numérica (CPN), pp. 56. SENAMHI, 2009d: Climate Scenarios for Peru to 2030. Available at: http://www.senamhi.gob.pe. National Meteorology and Hydrology Service (SENAMHI); Numerical Prediction Center (CPN), pp. 136. Seneviratne, S.I., M. Nicholls, D. Easterling, C.M. Goodess, S. Kanae, J. Kossin, Y. Luo, J.A. Marengo, K. McInnes, M. Rahimi, M. Reichstein, A. Sorteberg, C. Vera, and X. Zhang, 2012: Changes in climate extremes and their impacts on the natural physical environment. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation. In: A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change. [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al. (eds.)]. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 109-230. Seo, S.N., B.A. McCarl, and R. Mendelsohn, 2010: From beef cattle to sheep under global warming? An analysis of adaptation by livestock species choice in South America. Ecological Economics, 69(12), 2486-2494. Seoane, R. and P. López, 2007: Assessing the effects of climate change on the hydrological regime of the Limay River basin. GeoJournal, 70(4), 251-256. Seth, A., M. Rojas, and S.A. Rauscher, 2010: CMIP3 projected changes in the annual cycle of the South American Monsoon. Climatic Change, 98(3-4), 331-357. Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Shepard, D.S., L. Coudeville, Y.A. Halasa, B. Zambrano, and G.H. Dayan, 2011: Economic impact of dengue illness in the Americas. American Journal of Tropical Medicine and Hygiene, 84(2), 200-207. Sherbinin, A.d., M. Levy, S. Adamo, K. MacManus, G. Yetman, V. Mara, L. Razafindrazay, B. Goodrich, T. Srebotnjak, C. Aichele, and L. Pistolesi, 2012: Migration and risk: net migration in marginal ecosystems and hazardous areas. Environmental Research Letters, 7(4), 045602. Shiogama, H., S. Emori, N. Hanasaki, M. Abe, Y. Masutomi, K. Takahashi, and T. Nozawa, 2011: Observational constraints indicate risk of drying in the Amazon basin. Nature Communications, 2(253). SICA, 2013: CRRH-Comité Regional de Recursos Hidráulicos [Regional Committee of Hydraulic Resources]. Available at: http://www.recursoshidricos.org/. Sietz, D., B. Untied, O. Walkenhorst, M.K.B. Lüdeke, G. Mertins, G. Petschel-Held, and H.J. Schellnhuber, 2006: Smallholder agriculture in Northeast Brazil: assessing heterogeneous human-environmental dynamics. Regional Environmental Change, 6(3), 132-146. Sietz, D., 2011: Dryland vulnerability Typical patterns and dynamics in support of vulnerability reduction efforts. Diss. Doctor rerum naturalium in Global Ecology, Faculty of Science University of Potsdam, Germany/ Potsdam Institute for Climate Impact Research Research Domain II - Climate Impacts and Vulnerabilities, Potsdam, Germany. Sietz, D., M.K.B. Luedeke, and C. Walther, 2011: Categorisation of typical vulnerability patterns in global drylands. Global Environmental Change, 21(2), 431-440. Sietz, D., S.E. Mamani Choque, and M.K.B. Lüdeke, 2012: Typical patterns of smallholder vulnerability to weather extremes with regard to food security in the Peruvian Altiplano. Regional Environmental Change, 12(3), 489- 505. Silva Dias, M.A.F., J. Dias, L. Carvalho, E. Freitas, and P.L. Silva Dias, 2012: Changes in extreme daily rainfall for Sao Paulo, Brazil. Climatic Change, 116(3-4), 705-722. Silva, A.G. and P. Azevedo, 2008: Índices de tendencias de Mudanças Climáticas no Estado da Bahia [Indices of climate change trends in the State of Bahia]. Engenheiria Ambiental, 5, 141-151. Silva, T.G.F., M.S.B. Moura, I.I.S. Sá, S. Zolnier, S.H.N. Turco, F. Justino, J.F.A. Carmo, and L.S.B. Souza, 2009: Impactos das mudanças climáticas na produçao leiteira do estado de Pernambuco: análise para os cenários B2 e A2 do IPCC [Impacts of climate change on regional variation of milk production in the Pernambuco State, Brazil: analysis for the A2 and B2 IPCC scenarios]. Revista Brasileira De Meteorologia, 24(4), 489-501. Silva, V.d.P.R., J.H.B.C. Campos, M.T. Silva, and P.V. Azevedo, 2010: Impact of global warming on cowpea bean cultivation in northeastern Brazil. Agricultural Water Management, 97(11), 1760-1768. Siqueira, M.F.d. and A.T. Peterson, 2003: Consequences of global climate change for geographic distributions of Cerrado species. Biota Neotropica, 3(2), 1-14. Sitch, S., C. Huntingford, N. Gedney, P.E. Levy, M. Lomas, S.L. Piao, R. Betts, P. Ciais, P. Cox, P. Friedlingstein, C.D. Jones, I.C. Prentice, and F.I. Woodward, 2008: Evaluation of the terrestrial carbon cycle, future plant geography and climate-carbon cycle feedbacks using five Dynamic Global Vegetation Models (DGVMs). Global Change Biology, 14(9), 2015-2039. Smil, V., 2000: Energy in the twentieth century: Resources, conversions, costs, uses, and consequences. Annual Review of Energy and the Environment, 25, 21-51. Smolka, M.O. and A.A. Larangeira, 2008: Informality and poverty in Latin American urban policies (Chapter 5) . In: The New Global Frontier: Urbanization, Poverty and Environment in the 21st Century. [Martine, G., G. McGranahan, M. Montgomery, and R. Fernández-Castilla (eds.)]. Earthscan, London, UK, pp. 99. Soares, W.R. and J.A. Marengo, 2009: Assessments of moisture fluxes east of the Andes in South America in a global warming scenario. International Journal of Climatology, 29(10), 1395-1414. Soares-Filho, B., P. Moutinho, D. Nepstad, A. Anderson, H. Rodrigues, R. Garcia, L. Dietzsch, F. Merry, M. Bowman, L. Hissa, R. Silvestrini, and C. Maretti, 2010: Role of Brazilian Amazon protected areas in climate change mitigation. Proceedings of the National Academy of Sciences of the United States of America, 107(24), 10821-10826. Soito, J.L.d.S. and M.A.V. Freitas, 2011: Amazon and the expansion of hydropower in Brazil: Vulnerability, impacts and possibilities for adaptation to global climate change. Renewable and Sustainable Energy Reviews, 15(6), 3165-3177. Soleri, D., D.A. Cleveland, G. Glasgow, S.H. Sweeney, F. Aragon Cuevas, M.R. Fuentes, and H. Rios L., 2008: Testing assumptions underlying economic research on transgenic food crops for Third World farmers: Evidence from Cuba, Guatemala and Mexico. Ecological Economics, 67(4), 667-682. Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sörensson, A.A., C.G. Menéndez, R. Ruscica, P. Alexander, P. Samuelsson, and U. Willén, 2010: Projected precipitation changes in South America: a dynamical downscaling within CLARIS. Meteorologische Zeitschrift, 19(4), 347-355. Soriano, M., K.A. Kainer, C.L. Staudhammer, and E. Soriano, 2012: Implementing multiple forest management in Brazil nut-rich community forests: Effects of logging on natural regeneration and forest disturbance. Forest Ecology and Management, 268, 92-102. Sortino-Rachou, A.M., M.P. Curado, and M.d.C. Cancela, 2011: Cutaneous melanoma in Latin America: a population-based descriptive study. Cadernos De Saúde Pública, 27(3), 565-572. Soruco, A., C. Vincent, B. Francou, and J.F. Gonzalez, 2009: Glacier decline between 1963 and 2006 in the Cordillera Real, Bolivia. Geophysical Research Letters, 36, L03502. Southgate, D., T. Haab, J. Lundine, and F. Rodríguez, 2010: Payments for environmental services and rural livelihood strategies in Ecuador and Guatemala. Environment and Development Economics, 15(1), 21-37. Souvignet, M., H. Gaese, L. Ribbe, N. Kretschmer, and R. Oyarzun, 2010: Statistical downscaling of precipitation and temperature in north-central Chile: an assessment of possible climate change impacts in an arid Andean watershed. Hydrological Sciences Journal-Journal Des Sciences Hydrologiques, 55(1), 41-57. Souvignet, M., R. Oyarzún, K.M.J. Verbist, H. Gaese, and J. Heinrich, 2012: Hydro-meteorological trends in semi- arid north-central Chile (29-32 °S): water resources implications for a fragile Andean region. Hydrological Sciences Journal-Journal Des Sciences Hydrologiques, 57(3), 479-495. Souza Filho, F.d.A.d. and C.M. Brown, 2009: Performance of water policy reforms under scarcity conditions: a case study in Northeast Brazil. Water Policy, 11(5), 553-568. Souza, M.N., E.C. Mantovani, A.G.d. Silva Júnior, J.J. Griffiti, and R.C. Delgado, 2010: Evaluation of hydrologic behavior of the Entre Ribeiros river basin, an affluent of Paracatu river, in climatic change scenario, with the use of the Stella software. Engenharia Na Agricultura, 18(4), 339. SSN, 2006: The SouthSouthNorth Capacity Building Module on Poverty Reduction: Approaches for achieving sustainable development and poverty reduction [SSN Capacity Building Team (ed.)]. SouthSouthNorth (SSN), , pp. 63. Stehr, A., P. Debels, J.L. Arumí, H. Alcayaga, and F. Romero, 2010: Modelación de la respuesta hidrológica al cambio climático: experiencias de dos cuencas de la zona centro-sur de Chile. [Modeling the hydrological response to climate change: experiences from two south-central Chilean watersheds]. Tecnología y Ciencias Del Agua, 1(4), 37-58. Strassburg, B.B.N., A. Kelly, A. Balmford, R.G. Davies, H.K. Gibbs, A. Lovett, L. Miles, C.D.L. Orme, J. Price, R.K. Turner, and A.S.L. Rodrigues, 2010: Global congruence of carbon storage and biodiversity in terrestrial ecosystems. Conservation Letters, 3(2), 98-105. Strelin, J. and R. Iturraspe, 2007: Recent evolution and mass balance of Cordón Martial glaciers, Cordillera Fueguina Oriental. Global and Planetary Change, 59(1-4), 17-26. Sverdlik, A., 2011: Ill-health and poverty: a literature review on health in informal settlements. Environment and Urbanization, 23(1), 123-155. Tacconi, L., 2012: Redefining payments for environmental services. Ecological Economics, 73, 29-36. Tada, M.S., R.P. Marques, E. Mesquita, R.C. Dalla Martha, J.A. Rodrigues, J.D. Costa, R.R. Pepelascov, T.H. Katsuragawa, and L.H. Pereira-da-Silva, 2007: Urban malaria in the Brazilian Western Amazon Region I: high prevalence of asymptomatic carriers in an urban riverside district is associated with a high level of clinical malaria. Mem Inst Oswaldo Cruz, 102(3), 263-269. Takasaki, Y., 2007: Dynamic household models of forest clearing under distinct land and labor market institutions: can agricultural policies reduce tropical deforestation? Environment and Development Economics, 12(3), 423- 443. Tapia-Conyer, R., J.F. Méndez-Galván, and H. Gallardo-Rincón, 2009: The growing burden of dengue in Latin America. Journal of Clinical Virology, 46(SUPPL. 2), 3-6. Team, V. and L. Manderson, 2011: Social and public health effects of climate change in the '40 South'. Wiley Interdisciplinary Reviews: Climate Change, 2(6), 902-918. Teixeira, E.I., G. Fischer, H. van Velthuizen, C. Walter, and F. Ewert, 2013: Global hot-spots of heat stress on agricultural crops due to climate change. Agricultural and Forest Meteorology, 170, 206-215. Thompson, L.G., E. Mosley-Thompson, H. Brecher, M. Davis, B. Leon, D. Les, P.N. Lin, T. Mashiotta, and K. Mountain, 2006: Abrupt tropical climate change: past and present. Proc Natl Acad Sci USA, 103(28), 10536- 10543. Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Thompson, L.G., E. Mosley-Thompson, M.E. Davis, and H.H. Brecher, 2011: Tropical glaciers, recorders and indicators of climate change, are disappearing globally. Annals of Glaciology, 52(59), 23-34. Tirado, M.C., R. Clarke, L.A. Jaykus, A. McQuatters-Gollop, and J.M. Frank, 2010: Climate change and food safety: A review. Food Research International, 43(7), 1745-1765. Todd, M.C., R.G. Taylor, T.J. Osborn, D.G. Kingston, N.W. Arnell, and S.N. Gosling, 2011: Uncertainty in climate change impacts on basin-scale freshwater resources preface to the special issue: the QUEST-GSI methodology and synthesis of results. Hydrology and Earth System Sciences, 15(3), 1035-1046. Tomei, J. and P. Upham, 2009: Argentinean soy-based biodiesel: An introduction to production and impacts. Energy Policy, 37(10), 3890-3898. Tompkins, E.L., M.C. Lemos, and E. Boyd, 2008: A less disastrous disaster: Managing response to climate-driven hazards in the Cayman Islands and NE Brazil. Global Environmental Change-Human and Policy Dimensions, 18(4), 736-745. Tormey, D., 2010: Managing the effects of accelerated glacial melting on volcanic collapse and debris flows: Planchon-Peteroa Volcano, Southern Andes. Global and Planetary Change, 74(2), 82-90. Torres, J.R. and J. Castro, 2007: The health and economic impact of dengue in Latin America. Cadernos De Saude Publica, 23(SUPPL. 1), 23-31. Tourre, Y.M., L. Jarlan, J.-. Lacaux, C.H. Rotela, and M. Lafaye, 2008: Spatio-temporal variability of NDVI- precipitation over southernmost South America: Possible linkages between climate signals and epidemics. Environmental Research Letters, 3(4), 044008. Travasso, M.I., G.O. Magrin, W.E. Baethgen, J.P. Castano, G.I.R. Rodriguez, J.L. Pires, A. Gimenez, G. Cunha, and M. Fernández, 2008: Maize and Soybean Cultivation in Southeastern South America: Adapting to Climate Change. Chapter 19. In: Climate Change and Adaptation. [Leary, N., J. Adejuwon, V. Barros, I. Burton, J. Kulkarni, and R. Lasco (eds.)]. , pp. 332-352. Travasso, M.I., G.O. Magrin, M.O. Grondona, and G.R. Rodriguez, 2009a: The use of SST and SOI anomalies as indicators of crop yield variability. International Journal of Climatology, 29(1), 23-29. Travasso, M.I., G.O. Magrin, G.R. Rodríguez, S. Solman, and M. Núnez, 2009b: Climate change impacts on regional maize yields and possible adaptation measures in Argentina. International Journal of Global Warming, 1(1-3), 201-213. Troin, M., C. Vallet-Coulomb, F. Sylvestre, and E. Piovano, 2010: Hydrological modelling of a closed lake (Laguna Mar Chiquita, Argentina) in the context of 20th century climatic changes. Journal of Hydrology, 393(3-4), 233- 244. Trombotto, D. and E. Borzotta, 2009: Indicators of present global warming through changes in active layer- thickness, estimation of thermal diffusivity and geomorphological observations in the Morenas Coloradas rockglacier, Central Andes of Mendoza, Argentina. Cold Regions Science and Technology, 55(3), 321-330. Tschakert, P. and K.A. Dietrich, 2010: Anticipatory Learning for Climate Change Adaptation and Resilience. Ecology and Society, 15(2), 11. Tucker, C.M., H. Eakin, and E.J. Castellanos, 2010: Perceptions of risk and adaptation: Coffee producers, market shocks, and extreme weather in Central America and Mexico. Global Environmental Change, 20(1), 23-32. UGHR, 2010: Inventario de Glaciares Cordillera Blanca. Ministerio de Agricultura del Perú, Autoridad Nacional del Agua, Direccion de conservación y Planeamiento de Recursos Hidricos; Unidad de Glaciologia y Recursos Hidricos (UGHR), Huaraz, Peru, pp. 81. UN, 2010: Millennium Development Goals Advances in Environmentally Sustainable Development in Latin America and the Caribbean. United Nations (UN), Santiago, Chile, pp. 218. UNDP, 2007: Climate shocks: risk and vulnerability in an unequal world (Chapter 2). In: Human Development Report 2007/8. Fighting climate change: Human solidarity in a divided world. United Nations Development Programme (UNDP), New York, USA, pp. 71-108. UNDP, 2010: Regional human development report for Latin America and Caribean 2010. Acting on the future: breaking the intergenerational transmission of inequality. United Nations Development Programme (UNDP), San José, Costa Rica, pp. 208. UNEP-ECLAC, 2010: Vital Climate Change Graphics for Latin America and the Caribbean. UNEP-ECLAC, Santiago, Chile, pp. 37. UNFCCC, 2012: Non-Annex I national communications. Available at: http://unfccc.int/national_reports/non- annex_i_natcom/items/2979.php (accessed on 4 November 2012). Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 UN-Habitat, 2011: Climate Change Mitigation Responses in Urban Areas (Chapter 5). In: Cities and Climate Change. Global Report on Human Settlements 2011/ United Nations Human Settlements Programme. Earthscan, London/ Washington D.C., pp. 91-128. Urcola, H.A., J.H. Elverdin, M.A. Mosciaro, C. Albaladejo, J.C. Manchado, and J.F. Giussepucci, 2010: Climate Change Impacts on Rural Societies: Stakeholders Perceptions and Adaptation Strategies in Buenos Aires, Argentina. Proceedings of Innovation and Sustainable Development in Agriculture and Food; ISDA 2010, 28.June - 01.July 2010, Montpellier, France, pp. 10. Urrutia, R.B., A. Lara, R. Villalba, D.A. Christie, C. Le Quesne, and A. Cuq, 2011: Multicentury tree ring reconstruction of annual streamflow for the Maule River watershed in south central Chile. Water Resources Research, 47(6). Valderrama-Ardila, C., N. Alexander, C. Ferro, H. Cadena, D. Marín, T.R. Holford, L.E. Munstermann, and C.B. Ocampo, 2010: Environmental risk factors for the incidence of American cutaneous leishmaniasis in a sub- andean zone of Colombia (Chaparral, Tolima). American Journal of Tropical Medicine and Hygiene, 82(2), 243-250. Valdivia, C., A. Seth, J.L. Gilles, M. Garcia, E. Jimenez, J. Cusicanqui, F. Navia, and E. Yucra, 2010: Adapting to Climate Change in Andean Ecosystems: Landscapes, Capitals, and Perceptions Shaping Rural Livelihood Strategies and Linking Knowledge Systems. Annals of the Association of American Geographers, 100(4), 818- 834. Valentine, J., J. Clifton-Brown, A. Hastings, P. Robson, G. Allison, and P. Smith, 2012: Food vs. fuel: the use of land for lignocellulosic next generation' energy crops that minimize competition with primary food production. Global Change Biology Bioenergy, 4(1), 1-19. Valverde, M.d.l.A., J.M. Ramírez, L.G.M.d. Oca, M.G.A. Goris, N. Ahmed, and R.A. Hartskeerl, 2008: Arenal, a new Leptospira serovar of serogroup Javanica, isolated from a patient in Costa Rica. Infection, Genetics and Evolution, 8(5), 529-533. Van der Meide, W.F., A.J. Jensema, R.A.E. Akrum, L.O.A. Sabajo, R.F.M. Lai A Fat, L. Lambregts, H.D.F.H. Schallig, M. Van Der Paardt, and W.R. Faber, 2008: Epidemiology of cutaneous leishmaniasis in Suriname: A study performed in 2006. American Journal of Tropical Medicine and Hygiene, 79(2), 192-197. Van Noordwijk, M., B. Leimona, R. Jindal, G.B. Villamor, M. Vardhan, S. Namirembe, D. Catacutan, J. Kerr, P.A. Minang, and T.P. Tomich, 2012: Payments for Environmental Services: Evolution Toward Efficient and Fair Incentives for Multifunctional Landscapes. Annual Review of Environment and Resources, 37, 389. Van Oel, P.R., M.S. Krol, A.Y. Hoekstra, and R.R. Taddei, 2010: Feedback mechanisms between water availability and water use in a semi-arid river basin: A spatially explicit multi-agent simulation approach. Environmental Modelling & Software, 25(4), 433-443. Vargas, W.M., G. Naumann, and J.L. Minetti, 2011: Dry spells in the River Plata Basin: an approximation of the diagnosis of droughts using daily data. Theoretical and Applied Climatology, 104(1-2), 159-173. Venema, H.D. and M. Cisse, 2004: Seeing the Light: Adapting to climate change with decentralized renewable energy in developing countries. International Institute for Sustainable Development (IISD), Winnipeg, Manitoba, Canada, pp. 174. Venencio, M.d.V. and N.O. García, 2011: Interannual variability and predictability of water table levels at Santa Fe Province (Argentina) within the climatic change context. Journal of Hydrology, 409(1-2), 62-70. Vergara, W., A. Deeb, A. Valencia, R. Bradley, B. Francou, A. Zarzar, A. Grünwaldt, and S. Haeussling, 2007: Economic impacts of rapid glacier retreat in the Andes. Eos Trans. AGU, 88(25), 261-264. Viana, V.M., 2008: Bolsa Floresta (Forest Conservation Allowance): an innovative mechanism to promote health in traditional communities in the Amazon. Estudos Avançados [Online], 22(64), 143-153. Vich, A.I.J., P.M. López, and M.C. Schumacher, 2007: Trend detection in the water regime of the main rivers of the Province of Mendoza, Argentina. GeoJournal, 70(4), 233-243. Vicuna, S., R. Garreaud, J. McPhee, F. Meza, and G. Donoso, 2010: Vulnerability and Adaptation to Climate Change in an Irrigated Agricultural Basin in Semi Arid Chile. [Potter, K.W. and Frevert, D.K. (eds.)]. American Society of Civil Engineers (ASCE), Madison, Wisconsin, USA, pp. 13-13. Vicuna, S., J. Gironás, F.J. Meza, M.L. Cruzat, M. Jelinek, E. Bustos, D. Poblete, and N. Bambach, 2013: Exploring possible connections between hydrological extreme events and climate change in central south Chile. (accepted 18 February 2013). Hydrological Sciences Journal, 58. Vicuna, S., R.D. Garreaud, and J. McPhee, 2011: Climate change impacts on the hydrology of a snowmelt driven basin in semiarid Chile. Climatic Change, 105(3-4), 469-488. Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Vicuna, S., J. McPhee, and R.D. Garreaud, 2012: Agriculture Vulnerability to Climate Change in a Snowmelt Driven Basin in Semiarid Chile. Journal of Water Resources Planning and Management, 138(5), 431-441. Viglizzo, E.F., F.C. Frank, L.V. Carreno, E.G. Jobbágy, H. Pereyra, J. Clatt, D. Pincén, and M.F. Ricard, 2011: Ecological and environmental footprint of 50 years of agricultural expansion in Argentina. Global Change Biology, 17(2), 959-973. Viglizzo, E.F., E.G. Jobbagy, L. Carreno, F.C. Frank, R. Aragon, L. De Oro, and V. Salvador, 2009: The dynamics of cultivation and floods in arable lands of Central Argentina. Hydrology and Earth System Sciences, 13(4), 491-502. Viglizzo, E.F. and F.C. Frank, 2006: Ecological interactions, feedbacks, thresholds and collapses in the Argentine Pampas in response to climate and farming during the last century. Quaternary International, 158, 122-126. Vignola, R., B. Locatelli, C. Martinez, and P. Imbach, 2009: Ecosystem-based adaptation to climate change: what role for policy-makers, society and scientists? Mitigation and Adaptation Strategies for Global Change, 14(8), 691-696. Villacís, M., 2008: Ressources en eau glaciaire dans les Andes d Equateur en relation avec les variations du climat: le cas du volcan Antisana. [Resources of water ice in the Andes of Ecuador in relation to climate variations: the case of Antisana volcano.]. Diss. PhD, Université Montpellier, Montpellier, 256 pp. Volpe, F.M., E.M. da Silva, T.N. dos Santos, and D.E.G. de Freitas, 2010: Further evidence of seasonality of mania in the tropics. Journal of Affective Disorders, 124(1-2), 178-182. von Braun, J., 2007: The World Food Situation. New Driving Forces and Required Actions. International Food Policy Research Institute, Washington,D.C., USA, pp. 18. Vuille, M., B. Francou, P. Wagnon, I. Juen, G. Kaser, B.G. Mark, and R.S. Bradley, 2008a: Climate change and tropical Andean glaciers: Past, present and future. Earth-Science Reviews, 89(3-4), 79-96. Vuille, M., G. Kaser, and I. Juen, 2008b: Glacier mass balance variability in the Cordillera Blanca, Peru and its relationship with climate and the large-scale circulation. Global and Planetary Change, 62(1-2), 14-28. Walter, L.C., H.T. Rosa, and N.A. Streck, 2010: Simulaçao do rendimento de graos de arroz irrigado em cenários de mudanças climáticas [Simulating grain yield of irrigated rice in climate change scenarios]. Pesquisa Agropecuaria Brasileira, 45(11), 1237-1245. Walther, G., 2010: Community and ecosystem responses to recent climate change. Royal Society Philosophical Transactions Biological Sciences, 365(1549), 2019-2024. Wang, B., J. Liu, H. Kim, P.J. Webster, and S. Yim, 2012: Recent change of the global monsoon precipitation (1979-2008). Climate Dynamics, 39(5 SI), 1123-1135. Wang, G., S. Sun, and R. Mei, 2011: Vegetation dynamics contributes to the multi-decadal variability of precipitation in the Amazon region. Geophysical Research Letters, 38, L19703. Warner, J. and M.T. Oré, 2006: El Nino platforms: participatory disaster response in Peru. Disasters, 30(1), 102- 117. Wassenaar, T., P. Gerber, P.H. Verburg, M. Rosales, M. Ibrahim, and H. Steinfeld, 2007: Projecting land use changes in the Neotropics: The geography of pasture expansion into forest. Global Environmental Change- Human and Policy Dimensions, 17(1), 86-104. Waycott, M., C.M. Duarte, T.J.B. Carruthers, R.J. Orth, W.C. Dennison, S. Olyarnik, A. Calladine, J.W. Fourqurean, K.L. Heck Jr., A.R. Hughes, G.A. Kendrick, W.J. Kenworthy, F.T. Short, and S.L. Williams, 2009: Accelerating loss of seagrasses across the globe threatens coastal ecosystems. Proceedings of the National Academy of Sciences of the United States of America, 106(30), 12377-12381. WB, 2009: Assessing the Potential Consequences of in America Climate Destabilization in Latin America. In: Latin America and Caribbean Region Sustainable Development Working Paper No. 32 [Vergara, W. (ed.)]. The World Bank, Latin America and the Caribbean Region, Sustainable Development Department (LCSSD), pp. 118. WB, 2012: The World Bank Data. Accessed on 6 May 2012. In: World Development Indicators, Urban Development, Urban population (% of total) and Population in the largest city (%of urban population). Available at: http://data.worldbank.org/topic/urban-development The World Bank. Winchester, L., 2008: Harmony and Dissonance between Human Settlements and the Environment in Latin America and the Caribbean (LC/W.204. Project Document N 204). United Nations, ECLAC, Santiago de Chile, Chile. Winchester, L. and R. Szalachman, 2009: The urban poor's vulnerability to the impacts of climate change in Latin America and the Caribbean - A policy agenda. In: 5th Urban Research Symposium "Cities and Climate Change: Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Responding to the Urgent Agenda". United Nations (UN), Economic Commission for Latin America and the Caribbean (ECLAC). Wright, S.J. and M.J. Samaniego, 2008: Historical, Demographic, and Economic Correlates of Land-Use Change in the Republic of Panama. Ecology and Society, 13(2), 17. Xu, Y., X. Gao, and F. Giorgi, 2009: Regional variability of climate change hot-spots in East Asia Science Press, co-published with Springer-Verlag GmbH, , pp. 783-792. Young, G., H. Zavala, J. Wandel, B. Smit, S. Salas, E. Jimenez, M. Fiebig, R. Espinoza, H. Diaz, and J. Cepeda, 2010: Vulnerability and adaptation in a dryland community of the Elqui Valley, Chile. Climatic Change, 98(1- 2), 245-276. Young, K.R. and J.K. Lipton, 2006: Adaptive Governance and Climate Change in the Tropical Highlands of Western South America. Climatic Change, 78(1), 63-102. Zagonari, F., 2010: Sustainable, Just, Equal, and Optimal Groundwater Management Strategies to Cope with Climate Change: Insights from Brazil. Water Resources Management, 24(13), 3731-3756. Zak, M.R., M. Cabido, D. Caceres, and S. Diaz, 2008: What Drives Accelerated Land Cover Change in Central Argentina? Synergistic Consequences of Climatic, Socioeconomic, and Technological Factors. Environmental Management, 42(2), 181-189. Zhang, X. and X. Cai, 2011: Climate change impacts on global agricultural land availability. Environmental Research Letters, 6(1), 8. Zhang, Y., R. Fu, H. Yu, Y. Qian, R. Dickinson, M.A.F. Silva Dias, P.L. da Silva Dias, and K. Fernandes, 2009: Impact of biomass burning aerosol on the monsoon circulation transition over Amazonia. Geophysical Research Letters, 36, L10814. Zullo, J.,Jr., H.S. Pinto, E.D. Assad, and A.M. Heuminski de Avila, 2011: Potential for growing Arabica coffee in the extreme south of Brazil in a warmer world. Climatic Change, 109(3-4), 535-548. Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 27-1: Regional observed changes in temperature, precipitation and climate extremes in various sectors of CA and SA. Additional information on changes in observed extremes can be found in the IPCC SREX (Seneviratne et al., 2012) and Chapter 2 IPCC WGI AR5 [2.4, 2.5, 2.6] Region Period Observed changes CA and Northern SA/Variable Precipitation in the NAMS (Englehart and Douglas, 2006) 1943-02 +0.94 mm/day/58 years Rainfall onset in NAMS (Grantz et al., 2007) 1948-04 -10 to -20 days/57 years Summertime Precipitation in NAMS (Anderson et al., 2010) 1931-00 +17.6 mm century 1 Rainfall extremes (P95) in NAMS (Cavazos et al., 2008) 1961-98 +1.3% decade 1 Cold days and nights in CA and Northern SA (Donat et al., 2013) 1951-10 Cold days: -1 day decade 1 , Cold nights: -2 day decade 1 Warm days and nigths in Northern SA ((Donat et al., 2013) 1951-10 Warm days: +2 to +4 day decade 1 , warm nights: +1 to +3 day decade 1 Heavy precipitation (R10) in Northern SA ((Donat et al., 2013) 1951-10 +1 to +2 day decade 1 , Consecutive dry days (CDD) in Northern SA (Donat et al., 2013) 1951-10 -2 day decade 1 , West Coast of SA SST and air temperatures off coast of Peru and Chile (15S-35S) (Falvey and Garreaud, 2009; 1960-10 -0.25 oC /decade, -0.7 oC /11 years for 2002-2012 Gutiérrez et al., 2011a; Gutiérrez et al., 2011b; Kosaka and Xie, 2013) Temperature, precipitation, cloud cover, and number of rainy days since the middle 1970´s off 1920-09 -1 oC /40 years, -1.6 mm/40 years, -2 octas/40 years, and -0.3 coast of Chile (18S-30S) (Schulz et al., 2012) days/40 years Wet days until 1970, increase after that, reduction in the precipitation rate in southern Chile 1900-07 -0.34% until 1970 and +0.37 after that, -0.12 % (37S-43S) (Quintana and Aceituno, 2012) Cold days and nights in all South American coast (Donat et al., 2013) 1951-10 Cold days: -1 days decade-1; cold nights: -2 days decade-1 Warm nights in all South American coast, warm days in the northern coast of South America, 1951-10 Warm night: -1 days decade-1; warm days: +3 days decade-1 ; warm days off the coast of Chile (Donat et al., 2013) warm days: -1 days decade-1 Warm nights in the coast of Chile (Dufek et al., 2008) 1961-90 +5 to +9%/31 years Dryness as estimated by the Palmer Drought Severity Index (PDSI) for most of the west coast 1950-08 -2 to -4 / 50 years of SA (Chile, Ecuador, Northern Chile) (Dai, 2011) Heavy precipitation (R95) in northern and central Chile (Dufek et al., 2008) 1961-90 -45 to -105 mm/31 years Temperature and precipitation in southern Chile (Vicuna et al., 2013) 1976-08 -45 to -105 mm/31 years SESA Mean annual air temperature in southern Brazil (Sansigolo and Kayano, 2010) 1913-06 +0.5 to +0.6 oC /decade, -31.4 - -47.6 mm/decade Frequency of cold days and nights, warm days in Argentina and Uruguay (Rusticucci and 1935-02 -1.2%/decade, -1%/decade/, +0.2%/decade Renom, 2008) Highest annual maximum temperature, lowest annual minimum air temperature in Argentina 1956-03 +0.8 C/47 years, +0.6C/47 years and Uruguay (Rusticucci and Tencer, 2008) Warm nights in Argentina and Uruguay and southern Brazil (Rusticucci, 2012) 1960-09 10-20%/41 years Warm nights in most of the region (Dufek et al., 2008) 1961-90 +7 to +9%/31 years Cold nights in most of the region (Dufek et al., 2008) 1961-90 -5 to -9%/31 years Cold days and nights in most of the region (Donat et al., 2013) 1951-10 warm nights: +3 days decade-1; warm days: +4 days decade-1 Warm days and nights in most of the region (Donat et al., 2013) 1951-10 Cold nights: -3 days decade-1; cold days: -3 days decade-1 Consecutive dry days (CDD) in the La Plata Basin countries (Argentina, Bolivia and Paraguay) 1961-90 +15 to +21 days/31 years, -21 to -27 days/31 years and decrease of CDD in SA South of 30 S (Dufek et al., 2008) Number of dry months during the warm season October-March in the Pampas region between 1904-00 From 2-3 months in 1904-1920 to 1-2 months from 1980-2000 Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 25S-40S (Barrucand et al., 2007) Moister conditions as estimated by the Palmer Drought Severity Index (PDSI) in most of 1950-08 0 to 4/50 years SESA (Dai, 2011) Rainfall trends in the Parana River Basin (Dai et al., 2009) 1948-08 +1.5 mm/day/50 years Number of days with precipitation above 10 mm (R10) in most of the region (Donat et al., 1951-10 +2 days/decade-1 2013) Heavy precipitation (R95) in most of the region (Donat et al., 2013) 1951-10 +1% decade-1 and -4 days decade --1 Heavy precipitation (R95) in most of the region (Dufek et al., 2008) 1961-90 +45 to +135 mm/31 years Heavy precipitation (R95) in the state of Sao Paulo (Dufek and Ambrizzi, 2008) 1950-99 +50 to +75 mm/40 years Consecutive dry days (CDD) in the state of Sao Paulo (Dufek and Ambrizzi, 2008) 1950-90 -25 to-50 days/40 years Lightning activity significantly with change in temperature in the state of Sao (Pinto and Pinto, 1951-06 +40% per 1oC for daily and monthly 2008; Pinto et al., 2013) time scales and approximately 30% per 1 oC for decadal timescale Number of days with rainfall above 20 mm in the city of Sao Paulo (Marengo et al., 2013; 2005-11 +5 to +8 days/11 years Silva Dias et al., 2012) Excess rainfall events duration after 1950 (Krepper and Zucarelli, 2010) 1901-03 + 21 months/53 years Dry events and events of extreme dryness from 1972 to 1996 (Vargas et al., 2011) 1972-96 -29 days/24 years Number of dry days in Argentina (Rivera et al., 2013) 1960-05 -2 to -4 days/decade Extreme daily rainfall in La Plata Basin (Penalba and Robledo, 2010) 1950-00 +33 to +60% increase in Spring, Summer and Autumn, -10 to - 25% decrease in winter Frequency of heavy rainfall in Argentina, Southern Brazil and Uruguay (Re and Barros, 2009) 1959-02 +50 to +150 mm/43 years Annual precipitation in the La Plata Basin (Doyle et al., 2012)(Doyle and Barros, 2011) 1960-05 +5 mm/year (Doyle and Barros, 2011) Andes Mean maximum temperature along the Andes, and increase in the number of frost days 1921-10 +0.10 to +0.12 oC /decade in 1921-2010, and +0.23-0.24 oC (Marengo et al., 2011b) /decade during 1976-2010; 8 days/decade during 196-2002 Air temperature and changes in precipitation Northern Andes (Colombia, Ecuador) (Villacís, 1961-90 +0.1 C to +0.22 oC /decade, -4 to +4 %/decade years 2008) Temperature and precipitation in northern and central Andes of Peru (SENAMHI, 2005; 2007; 1963-06 +0.2 to +0.45 oC /decade, -20 to -30%/40 years 2009a; 2009c; 2009d) Temperature and precipitation in the southern Andes of Peru (Marengo et al., 2011b; 1964-06 +0.2 to 0.6 oC /decade, -11 to +2 mm/decade SENAMHI, 2007; 2009a; 2009b; 2009c; 2009d) Air temperature and rainfall over Argentinean and Chilean Andes and Patagonia (Falvey and 1950-90 +0.2 to 0.45 oC /decade, -10 to -12%/decade Garreaud, 2009; Masiokas et al., 2008) Number of days with rainfall above 10 mm (R10) (Donat et al., 2013) 1950-10 -3 days decade-1 Dryness in the Andes between 35.65 S-39.9 S using the PDSI (Christie et al., 2011) 1950-03 -7 PDSI/53 years Rainfall decrease in the Mantaro Valley, central Andes of Peru (SENAMHI, 2009c) 1970-05 -44 mm/decade Air temperature in Colombian Andes (Poveda and Pineda, 2009) 1959-07 +1 oC /20 years Amazon region Decadal variability of rainfall in northern and southern Amazonia (Marengo et al., 2009; 1920-08 -3 STD/30 years in northern Amazonia and +4 STD/30 years in Satyamurty et al., 2010) southern Amazonia since the middle 1970 s Rainfall in all the region (Espinoza et al., 2009a; 2009b) 1975-03 -0.32 %/28 years Onset of the rainy season in southern Amazonia (Butt et al., 2011; Marengo et al., 2011b) 1950-10 -1 month since 1976 to 2010 Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Precipitation in the SAMS core region (Wang et al., 2012) 1979-08 + 2 mm/day decade 1 Onset becomes steadily earlier from 1948 to early 1970s, demise dates have remained later, 1948-08 SAMS from 170 days (1948 1972) to 195 days (1972 1982). and SAMS duration was longer after 1972 (Carvalho et al., 2011) Spatially varying trends of heavy precipitation (R95), increase in many areas and insufficient 1961-90 +100 mm/31 years in western and extreme eastern Amazonia, evidence in others (Marengo et al., 2009) Spatially varying trends in dry spells in (CDD), increase in many areas and decrease in others 1961-90 +15 mm/31 years in western Amazonia, -20 mm/ in southern (Marengo et al., 2009; Marengo et al., 2010) Amazonia Rainfall in most of Amazonia and in western Amazonia (Dai et al., 2009; Dai, 2011) 1948-08 +1 mm/day/50 years, -1.5 mm/day/50 years Dryness as estimated by the Palmer Drought Severity Index PDI in southern Amazonia and 1950-08 -2 to -4/50 years, +2 to +4 /50 years moister conditions in western Amazonia (Dai, 2011) Seasonal mean convection and cloudiness (Arias et al., 2011) 1984-07 +30 W/m2/23 years, -8 %/23 years Onset of rainy season in southern Amazonia due to land use change (Butt et al., 2011) 1970-10 -0.6 days/30 years Precipitation in the region (Gloor et al., 2013) 1990-10 -20 mm/21 years Northeast Brazil Rainfall trends interior Northeast Brazil and in northern Northeast Brazil (Dai et al., 2009; Dai, 1948-08 -0.3 mm/day/50 years, +1.5 mm/day/50 years 2011) Heavy precipitation (R95) in some areas, and in southern Northeast Brazil (Silva and Azevedo, 1970-06 -2 mm/24 years to + 6 mm/24 years, 2008) Consecutive dry days CDD in most of southern Northeast Brazil (Silva and Azevedo, 2008) 1970-06 -0.99 days/24 years Total annual precipitation in northern Northeast Brazil (Santos and Brito, 2007) 1970-06 +1 to +4 mm/year/24 years Spatially varying trends in heavy precipitation (R95) in northern Northeast Brazil (Santos and 1970-06 -0.1 to +5 mm/yeas/24 years Brito, 2007) Spatially varying trends in heavy precipitation (R95) and consecutive dry days (CDD) in 1935-06 -0.4 to +2.5 mm/year/69 years, -1.5 to +1.5 days/year/69 years, northern Northeast Brazil (Santos et al., 2009) Dryness in Southern Northeast Brazil as estimated by the PDSI, and northern Northeast Brazil 1950-08 -2 to -4/50 years, 0 to +1/50 years (Dai, 2011) Subject to Final Copyedit 85 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 27-2: Regional projected changes in temperature, precipitation, and climate extremes in different sectors of CA and SA. Various studies used A2 and B2 scenarios from CMIP3 and various RCPs scenarios for CMIP5, and different time slices from 2010 to 2100. In order to make results comparable, the CMIP3 and CMIP5 at the time slice ending in 2100 are included. Additional information on changes in projected extremes can be found in the IPCC SREX (see IPCC, 2012), and Chapters 9 and 14 from IPCC WG1AR5 [9.5, 9.6 and 14.2, 14.7] Region Models and scenarios Projected changes CA and Northern SA Leaf Area Index (LAI), evapotranspiration by 2070-2099 in CA (Imbach et al., 2012) 23 CMIP3 models, A2 Evapotranspiration: +20%; LAI:-20%+0.94 mm/day/58 years Air temperature by 2075 and 2100 in CA (Aguilar et al., 2009) 9 CMIP3 models, A2 +2.2 C by 2075; +3.3 C by 2100 Rainfall in CA and Venezuela, air temperature in the region (Hall et al., 2013; Kitoh et al., 2011) 20 km MRI-AGCM3.1S Rainfall decrease/increase of about - model, A1B 10%/+10%, by 2079. Temperature increases of about +2.5-+3.5 oC by 2079 Precipitation and evaporation in most of the region. Soil moisture in most land areas in all 20 km MRI-AGCM3.1S Precipitation decrease of about -5 mm/day, seasons (Nakaegawa et al., 2013b) model, A1B evaporation increase of about +3 to +5 mm/day; soil moisture to decrease by -5 mm/day. Rainfall in Nicaragua, Honduras, Northern Colombia and Northern Venezuela, rainfall in Costa PRECIS forced by the Rainfall: -25 to -50%, and +25 to +50%. Rica and Panama. Temperature in all region by 2071-2100 (Campbell et al., 2011) HadAM3, A2 Temperature: +3 to +6 C Precipitation and temperature in northern SA, decrease in interior Venezuela, temperature Eta forced with HadCM3, Increases by +30 to 50%, and reductions increases by 2071-2100 (Marengo et al., 2011a) A1B between -10 to -20%; temperature: +4 to +5 C Precipitation and temperature by 2100 in CA (Karmalkar et al., 2011) PRECIS forced with Precipitation: -24 to -48%; temperature: +4 to HadAM3, A2 +5 C Warm nights, consecutive dry days and heavy precipitation in Venezuela, by 2100 (Marengo et PRECIS forced with Increase by +12 to +18%, +15 to +25 days and al., 2009; 2010) HadAM3, A2 reduction of 75 to 105 days Air temperature and precipitation in CA by 2100 (Giorgi and Diffenbaugh, 2008) 23 CMIP3 models, A1B Increase by +3 to +5 oC; reduction by -10 to - 30% Consecutive dry days and in heavy precipitation by 2099 (Kamiguchi et al., 2006) 20 km MRI-AGCM3.1S Increase by +5 days, and between +2 to +8 % model, A1B Rainfall over Panama by 2099 (Fábrega et al., 2013) 20 km MRI-AGCM3.1S Increase by +5 % model, A1B West Coast of SA Precipitation, runoff and temperature at the Limari river basin in semi-arid Chile by 2100 PRECIS forced with Precipitation: -15 % to -25%; runoff: -6 to - (Vicuna et al., 2011) HadAM3, A2 27%; temperature: + 3 to +4 oC Ai temperature and surface winds in west coast of SA (Chile) by 2100 (Garreaud and Falvey, 15 CMIP3 models, Temperature: +1 oC; coastal winds: +1.5 m/sec 2009) PRECIS forced with HadAM3, A2 Precipitation in the bands 5N-10S, and 25S-30S, and 10S-25S and 30S-50S; temperature Eta model forced with Increases of 30-40%; increases of 3 to 5 oC increase between by 2100 (Marengo et al., 2011a) HadCM3, A1B Warm nights, consecutive dry days, and heavy precipitation in 5N-5S by 2100 (Marengo et al., PRECIS forced with Increase of +3 to +18%, reduction of -5 to 8 2009; 2010) HadAM3, A2 days, increase by +75 to +105 days Air temperature, increase of precipitation between 0 and 10S, and between 20 and 40S by 2100 23 CMIP3 models, A1B Increase of +2 to +3 oC; increase by 10%, (Giorgi and Diffenbaugh, 2008) reduction by -10 to -30% Subject to Final Copyedit 86 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Consecutive dry days between 5 N and 10 S and south of 30S, heavy precipitation between 5S- 20 km MRI-AGCM3.1S Increase by 10 days and between +2 to +10% 20S and south of 20S by 2099 (Kamiguchi et al., 2006) model, A1B Precipitation between 15 and 35 S; and south of 40S; temperature by 2100 (Nunez et al., 2009) MM5 forced with Precipitation: -2 mm/day; +2 mm/day; HadAM3, A2 temperature: +2.5 oC Precipitation in Panama and Venezuela, heavy precipitation in Panama and in Venezuela, RCA forced with Precipitation:-1 to -3 mm/day, consecutive dry days over Panama and Colombia by 2099 (Sörensson et al., 2010) ECHAM5-MPI OM model, A1B SESA Precipitation and runoff, an in air temperature by 2100 (Marengo et al., 2011a) Eta forced with HadCM3, Precipitation: + 20 to +30%; Runoff: +10 to A1B +20%; air temperature: +2.5 to +3.5 oC Precipitation and temperature in the La Plata basin by 2050 (Cabré et al., 2010) MM5 forced with Precipitation: +0.5 to 1.5 mm/day; HadAM3, A2 temperature: +1.5 oC to 2.5 oC. Warm nights, consecutive dry days and heavy precipitation by 2100 (Menendez and Carril, 7 CMIP3 models, A1B Warm nights: +10 to +30%; Consecutive dry 2010) days: +1 to +5 days; Heavy precipitation: +3 to +9 %. Precipitation during summer and spring, and in fall and winter by 2100 (Seth et al., 2010) 9 CMIP3 models, A2 Precipitation: + 0.4 to +0.6 mm/day, -0.02 to - 0.04 mm/day Warm nights, consecutive dry days and heavy precipitation by 2100 (Marengo et al., 2009; PRECIS forced with Increase of +6 to +12%, +5 to +20 days, +75 2010) HadAM3, A2 to +105 days Air temperature and rainfall by 2100 (Giorgi and Diffenbaugh, 2008) 23 CMIP3 models, A1B Increase by +2 to +4 oC, increase by +20 to +30 % Consecutive dry days and in heavy precipitation by 2099 (Kamiguchi et al., 2006) 20 km MRI-AGCM3.1S Increase by +5 to +10% and by +2 to +8 % model, A1B Precipitation in north central Argentina, decrease in southern Brazil, increase of air temperature MM5 forced with Increase of +0.5 to +1 mm/day, reduction of - by 2100 (Nunez et al., 2009) HadAM3, A2 0.5 mm/day, increase of +3 to +4.5 oC Drought frequency, intensity and duration in South America South of 20 S for 2011-40 relative 15 CMIP5 models, RCP Frequency increase between 10-20%, increase to 1979-2008 (Penalba and Rivera, 2013) 4.5 and 8.5 in severity between 5-15% and reduction in duration between 10 and 30% Precipitation, heavy precipitation, reduction of consecutive dry days in the eastern part of the RCA forced with the Increase of +2 mm/day, of +5 to +15 mm, region, increase in the western part of the region by 2099 (Sörensson et al., 2010) ECHAM5 mode, A1B reduction of -10 days and increase of +5 days Precipitation in SESA by 2100 (Sörensson et al., 2010) 9 CMIP3 models, A1B Increase between +0.3 to +0.5 mm/day Andes Precipitation and temperature, increase by 2100 in the Altiplano (Minvielle and Garreaud, 2011) 11 CMP3 models, A2 Precipitation: -10 to -30 %; temperature:>3 oC Precipitation at 5N-5S, and 30S-45 S, at 5-25 S; temperature by 2100 (Marengo et al., 2011a) Eta forced with HadCM3, Increase between +10 and +30%, decrease by A1B -20 to -30%, increase of +3.5 to +4.5 oC Warm nights, heavy precipitation and consecutive dry days south of 15 S by 2100 (Marengo et PRECIS forced with Increase by +3 to +18%, reduction by -10 to - al., 2009) HadAM3, A2 20 days, and -75 to -105 days Air temperature, rainfall between 0-10S and reduction between 10-40 S (Giorgi and 23 CMIP3 models, A1B Increase by +3 to +4 oC, increase by 10% and Diffenbaugh, 2008) reduction by -10% Consecutive dry days and increase of heavy precipitation by 2099 (Kamiguchi et al., 2006) 20 km MRI-AGCM3.1S Reduction by -5 days, increase by +2 to +4 % model, A1B south of 20S Precipitation, heavy precipitation, and consecutive dry days by 2070-99 (Sörensson et al., 2010) RCA forced with Increases of +1 to +3 mm/day, +5 mm and of ECHAM5, A1B +5 to +10 days Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Summer precipitation and surface air temperature in the Altiplano region by 2099 (Minvielle and 9 CMIP3 models, A2 Reduction in precipitation between -10% and - Garreaud, 2011) 30%, an temperature increase of +3 oC Temperature and rainfall in lowland Bolivia inn 2070-2999 (Seiler et al., 2013) 5 CMIP3 models (A1B) Increase from 2.5 oC to 5 oC, reduction of 9% ad 5 CMIP5 models annual precipitation (RCP4.5, 8.5) Precipitation in the dry season, temperature and evapotranspiration 2079-98 (Guimberteau et al., CMIP3 models, A1B -1.1 mm; +2 oC; +7% 2013) Amazon region Rainfall in central and eastern Amazonia, and in western Amazonia, air temperature in all region Eta forced with HadCM3, Precipitation: -20 to -30%, +20 to +30%; by 2100 (Marengo et al., 2011a) A1B temperature: +5 to +7 oC Intensity of the South Atlantic Convergence Zone and in rainfall in the South American 10 CMIP3 models, A1B Precipitation: -100 to -200 mm/20 years monsoon region, 2081-2100 (Bombardi and Carvalho, 2009) Precipitation in western Amazonia during summer and in winter in Amazonia by 2100 (Mendes 5 CMIP3 models, A2 and +1.6% in summer and -1.5% in winter and Marengo, 2010) ANN Number of South American Low Level Jet east of the Andes events (SALLJ), and in the PRECIS forced by +50 events of SALLJ during summer, moisture transport from Amazonia to the La Plata basin by 2090 (Soares and Marengo, 2009) HadAM3, A2 increase in moisture transport by 50% Precipitation in the South American monsoon during summer and spring, and during fall and 9 CMIP3 models, A2 Increase of +0.15 to +0.4 mm/, reductions of - winter by 2100 (Seth et al., 2010) 0.10 to -0.26 mm/day Warm nights, consecutive dry days in eastern Amazonia, heavy precipitation in western PRECIS forced with Increase of +12 to +15%, by 25-30 days in Amazonia and in eastern Amazonia by 2100 (Marengo et al., 2009) HadAM3, A2 eastern Amazonia, increase in western Amazonia by 75-105 days and reduction by - 15 to 75 days in eastern Amazonia Increase in air temperature, rainfall increase in western Amazonia and decrease in eastern CMIP3 models, A1B Increase of +4 to +6 oC, increase of +10% and Amazonia by 2100 (Giorgi and Diffenbaugh, 2008) decrease between -10 to -30% Reduction of consecutive dry days and increase in heavy precipitation by 2099 (Kamiguchi et 20 km MRI-AGCM3.1S Reduction of -5 to -10 days, increase by +2 to al., 2006) model, A1B +8 % Onset and late demise of the rainy season in SAMS by 2040-2050 relative to 1951-80 (Jones and 10 CMIP5 models, Onset 14 days earlier than present, demise 17 Carvalho, 2013) RCP8.5 days later than present Precipitation in SAMS during the monsoon wet season in 2071-2100 relative to 1951-80 (Jones 10 CMIP5 models, Increase of 300 mm during the wet season and Carvalho, 2013) RCP8.5 Precipitation in western Amazonia, heavy precipitation in northern Amazonia and in southern RCA forced with the Increase of +1 to +3 mm/day, reduction of -1 Amazonia, consecutive dry days in western Amazonia and increase by 2099 (Sörensson et al., ECHAM5 model, A1B to -3 mm, increase of +5 to +10 mm, decrease 2010) of -5 to -10 days, increase by +20 to +30 days Northeast Brazil Rainfall and temperature in the entire region by 2100 (Marengo et al., 2011a) Eta forced with HadCM3, Precipitation: -20 to -20%; temperature: +3 to A1B +4 oC Warm nights, consecutive dry days, heavy precipitation by 2100 (Marengo et al., 2009) PRECIS forced with Increase by +18 to +24%, by +25 to +30 days HadAM3, A2 and -15 to 75 days Air temperature and precipitation by 2100 (Giorgi and Diffenbaugh, 2008) 23 CMIP3 models, A1B Increase of +2 to +4 oC, reduction -10 to -30% Consecutive dry days and heavy precipitation by 2099 (Kamiguchi et al., 2006) 20 km MRI-AGCM3.1S Reduction of -5 to -10% and increase of +2 to model, A1B +6 % Precipitation, heavy precipitation and consecutive dry days by 2099 (Sörensson et al., 2010) RCA forced with Increase of +1 to +2 mm/day, increase by +5 ECHAM5 model, A1B to +10 mm, and increase by +10 to +30 days Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 27-3: Observed trends related to Andean cryosphere. a) Andean tropical glacier trends Country Significant changes recorded References Documented massifs (latitude) (Variable code numbera) Value of trend (period of observed trend) Venezuela (1) +300/+500m between Little Ice Age (LIA) maximum and today. Polissar et al. (2006); Morris et al. (2006) Cordillera de Merida (10°N) (5) Accelerated melting since 1972. Risk of disappearing completely as ELA close to the highest peak (Pico Bolívar, 4979 m) Colombia (3) LIA maximum between AD 1600 and 1850 Ruiz et al. (2008); Ceballos et al. (2006); Parque Los Nevados (4°50N) (3) Many small/low elevation (<5000 m a.s.l.) glaciers have disappeared Poveda and Pineda (2009); IDEAM Sierra Nevada del Cocuy (6°30N) (3) -60/-84% (1850-2000); -50% (last 50 yrs); -10/-50% (past 15yrs); Retreat 3.0km/yr (2012); Rabatel et al. (2013) Sierra Nevada de Santa Marta (since 2000) (10°40N) Ecuador (1) +300 between the middle 18th (maximum LIA) and the last decades of the 20th C; Francou et al. (2007); Vuille et al. (2008); Antisana (0°28S) ~+200m (20th C) Jomelli et al. (2009); Cáceres (2010); Chimborazo and Carihuayrazo (3) ~-45% (1976-2006). Glaciers below 5300m in process of extinction Rabatel et al. (2013) (1°S) Peru (1) ~+100m (between LIA maximum and beginning 20th C); +150m (20th C) Raup et al. (2007); Jomelli et al. (2009); Cordillera Blanca (9°S) (3) -12/-17% (18th C); -17/-20% (19th C); -20/-35% (1960s-2000s); -26% (1962 and 2000) UGHR (2010); Bury et al. (2011); Mark et (4) -8 m decade 1 since 1970 (Yanamarey glacier) al. (2010); Baraer et al. (2012); Rabatel et (8) +1.6% (+/- 1.1) (glaciers with >20 percent glacier area); (7) Seven out of nine watersheds al. (2013) decreasing dry-season discharge Coropuna volcano (15°33S) (3) -26% (1962-2000) Racoviteanu et al. (2007) Cordillera Vilcanota (13°55S) (3) 10 times faster in 1991-2005 compared to 1963-2005 Thompson et al. (2006; 2011) (2-4) about -30% of area and about -45% of volume since 1985 Salzmann et al. (2013) Bolivia (1) +300 m (between LIA maximum and late 20th C); +180/+200m (20th C) Rabatel et al. (2006; 2008); Francou et al. Cordillera Real and Cordillera (2007); Vuille et al. (2008); Jomelli et al. (3) -48% (1976-2006) in the Cordillera Real; Chacaltaya vanished in 2010. Quimza Cruz (16°S) (2011); Soruco et al. (2009); Gilbert et al. (5) Zongo glacier have lost a mean of 0.4m w.e./yr in the 1991-2011 period; glaciers in the (2010); Rabatel et al. (2013) Cordillera Real lost -43% of their volume from 1963 to 2006 (maximum from 1976 to 2006). (2) +1.1+/-0.2°C over the 20th century at ~6340 m a.s.l. Sur Lipez, Caquella, 21°30S (7) Evidence of recent degradation of Caquella rock glacier (South Bolivian Altiplano) Francou et al. (1999) b) Extra tropical Andean cryosphere (glaciers, snowpack, runoff effects) trends. Region Significant changes recorded and reference References Documented massifs/latitude (Variable code numbera) Value of trend (period of observed trend) Chile, Argentina and Bolivia and (6) No significant trend. Foster et al. (2009) Argentinean Patagonia Dessert Andes (17°S-31°S) Subject to Final Copyedit 89 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Huasco basin glaciers (29°S) (5) 0.84 m w.e./yr (2003/2004 2007/2008) Nicholson et al. (2009); Rabatel et al. (2011); Gascoin et al. (2011) Central Andes (31°S-36°S) Piloto/Las Cuevas (32°S) (5) - 10.50 m w.e. (last 24 yrs) Leiva et al. (2007) Aconcagua basin glaciers (33°S) (3) -20% (last 48 years); (2) -14% (1955-2006). Nicholson et al. (2009); Bown et al. (8) Significant decrease in Aconcagua basin streamflow (2008); Pellicciotti et al. (2007) Central Andes glaciers (33 36 °S) (3) -3% (since 1955); (4) 50/ 9my 1 (during 20th C); (5) -0.76/-0.56 m/y (during 20th C) Le Quesne et al. (2009) ELA across central Andes (1) +122 +/- 8 m (winter) and 200 +/- 6 m (summer) (1975 and 2001) Carrasco et al. (2005) Snowpack (30 °S -37°S) (6) Positive, though nonsignificant, linear trend (1951 2005); (8) Mendoza river Masiokas et al. (2006) streamflow. Possible link to rising temperatures and snowpack/glacier effects. Not Vich et al. (2007) conclusive; (8) Increase in high and low flows possibly associated with increase in Vicuna et al. (2013) temperature and effects on snowpack Morenas coloradas rock glacier (7) Significant change in active layer possibly associated with warming processes. Trombotto and Borzotta (2009) (32-33°S) Cryosphere in the Andes of (5) Expansion of thermokarst depressions Bodin et al. (2010) Santiago (33.5°S.) Basins (28-47 °S) (8) Not significant increase in February runoff possible increase of glacier melt (1950 Casassa et al. (2009) 2007). Basins (30-40 °S) (8) Significant negative timing trend (CT date shifting towards earlier in the year) for 23 Cortés et al. (2011) out of the 40 analyzed series. Patagonian Andes (36°S-55°S) Streamflow from basins (28-47 °S) (8) Not significant increase in February run-off trends that might suggest an increase of Casassa et al. (2009) glacier melt in the Andes (1950 2007). NW Patagonia (38°-45°S) (4) Recession of 6 glaciers based on areal photograph analysis. Masiokas et al. (2008) Casa Pangue glacier (41°S) (5) 2.3+/-0.6 m/y (1961-1998); (4) 3.6+/-0.6 m/y (1981-1998) Bown and Rivera (2007) Manso  Glacier  (41°S) (8) Reduction in discharge associated with reduction in melt and precipitation Pasquini et al. (2013) North Patagonian Icefield (NPI) (8) Glacial lake outburst flood (GLOF) possible response to retreat of Calafate glacier (20th Harrison et al. (2006) C) Southern Patagonia Icefield (SPI) (4) Larger retreating rates observed on the west side coinciding with lower elevations of the Barcaza et al. (2009) ELAs NPI, SPI, Cordillera Darwin (4) 5.7-12.2 km (1945-2005) Lopez et al. (2010) Icefield Gran Campo Nevado (GCN) (53 (5) Slow retreat from Late LIA. Acceleration started 60 years ago Strelin and Iturraspe (2007) °S) (5) -2.8% of glacier length per decade (1942-2002); (3) -2.4% per decade (1942-2002) Schneider et al. (2007) Cordón Martial glaciers (54 °S) (5) -1.6 m/year or -27.9 +/- 11 km3/year (2002-2006) Chen et al. (2007) Proglacial lakes (40-50°S) (8) Summertime negative trend on lakes indicating that melt water is decreasing Pasquini et al. (2008) Notes: a Variable coding: (1) Increase in Equilibrium line altitude (ELA); (2) Atmospheric warming revealed by englacial temperature measured at high elevation; (3) Area reduction; (4) Frontal retreat; (5) Volume (water equivalent) reduction; (6) Snow cover; (7) Rock glaciers; (8) Runoff change; Subject to Final Copyedit 90 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 27-4: Synthesis of projected climate change impacts on hydrologic related variables in CA and SA basins and major glaciers. Notes: a Variable coding: (1) Runoff/Discharge; (2) Demand; (3) Recharge; (4) Glaciar change; (5) Unmet demand/water availability Region Hydrologic Variable Projected Changea Period GCM (GHG Scenarios) References Basins studied La Plata Basin and SESA Paraná (1): + 4.9% (not robust) 2081-2100 CMIP3 (A1B) Nohara et al. (2006) (1): +10 to +20% 2100 Eta-HadCM3 (A1B) Marengo et al. (2011a) (1) + 18.4% (significant) 2075-2100 CMIP3 models (A1B) Nakaegawa et al. (2013a) Grande (Parana) (1) +20/-20% Different periods 7 CMIP3 models Todd et al. (2011) ; Gosling et al. (2011); Nóbrega et al. (2011) Itaipu (Parana) (1) 2010 2040: Left bank: 5/ 15%; Right 2010 2040 and CCCMA-CGCM2 (A2) Rivarola et al. (2011) bank: +30%; (1) 2070-2100: 0 to 30% 2070-2100 Concordia (1): -40% 2070-2100 HadRM3P Perazzoli et al. (2013) Carcaraná (2) Increase 2010-2030 HadCM3 (A2) Venencio and García (2011) (3) Slight reduction Amazon Basin Peruvian Amazon (1) Some basins increased, some reduced Three time slices BCM2, CSMK3 and MIHR (A1B, Lavado et al. (2011) B1) Alto Beni-Bolivia (1) Increase and reduction. 2070-2100 CMIP3 models (A1B) Fry et al. (2012) (3) Always reduction (5) Increase in water stress Ecuador - (1) Some scenarios increase some reduction 2070-2100 CMIP3 models (A1B) Buytaert et al. (2011) Tomebamba/Paute Amazon at Obidos (1) + 5.4% (not robust) 2081-2100 CMIP3 models (A1B) Nohara et al. (2006) (1) +6% 2000-2100 ECBilt-CLIO-VECODE (A2) Aerts et al. (2006) (1) + 3.7% (significant) 2075-2100 CMIP3 models (A1B) Nakaegawa et al. (2013a) (1) No change in high flow. Reduction in low 2046 65/2079 98 8 AR4 GCMs (B1, A1B and A2) Guimberteau et al. (2013) flow Amazon -Orinoco (1) -20% 2050s HadCM3 (A2) Palmer et al.(2008) Brazil (1) Consistent decrease 2050s HadCM3 (A1b) and CMIP3 Arnel and Gosling (2013) Tropical Andes Colombian (4) Disappearance by 2020s linear extrapolation Poveda and Pineda (2009) glaciers Cordillera Blanca (1) Increase for next 20-50 years, reduction 2005-2020 Temperature output only (B2) Chevallier et al. (2011) glaciarized basins afterwards (4) 2050: area -38/-60%. Increased seasonality 2050 (climatology) Not specified (A1, A2, B1, B2) Juen et al. (2007) (4) 2080: area -49/-75%. Increased seasonality (4) Increased seasonality 2030 16 CMIP3 (A1B, B1) Condom et al. (2012) Basins providing (5) Inner tropics: Only small change. Increase 2010-2039 and 2040- 19 CMIP3 models (A1B, A2) Buytaert and De Bievre (2012) water to cities of in precipitation an increase in ET. 2069 Bogota, Quito, (5) Outer tropics: Severe reductions. Decrease Lima and La Paz in precipitation and increase in ET Subject to Final Copyedit 91 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Central Andes Limari (1) -20/-40%. 2070-2100 HadCM3 (A2, B2) Vicuna et al. (2011) (1) -20% 2010-2040 15 CMIP3 (A1B, B2, B1) Vicuna et al. (2012) (1) -30/-40%; (1) Change in seasonality 2070-2100 Maipo (1) -30% Three 30-year periods HadCM3 (A2, B2) Melo et al. (2010); ECLAC (2009); Meza (5) Unmet demand up to 50% 2070-2090 et al. (2012) Mataquito (1) Reduction in average and low flows. Three 30-year periods CMIP3 (A2, B1) and CMIP5 Demaria et al. (2013) Increase in high flows (RCP4.5 and 8.5) Maule, Laja (1) -30% Three 30-year periods HadCM3 (A2, B2) McPhee et al. (2010); ECLAC (2009) Bio Bio (1) (-81%/+ 7%) 2070-2100 8 GCMs (6 SRES) Stehr et al. (2010) Limay (1) -10/-20%. 2080s (climatology) HadCM2 (NS) Seoane and López (2007) North East Brazil (NEB) Brazilian Federal(1) No significant change up to 2025. (1) 2000-2100 HadCM2, ECHAM4 (NS) Krol et al. (2006); Krol and Bronstert States of Ceara´ After 2025: strong reduction with ECHAM4; (2007) and Piau ´ slight increase with HadCM2. Paracatu (Sao (1) A2: +31/+131%; 2000-2100 HadCM3 (A2, B2) De Mello et al. (2008) Francisco) (1) B2: no significant change Jaguaribe (2) Demand: +33/+44% 2040 HadCM3 (A2, B2) Gondim et al. (2012; 2008)(2008) (2) Irrigation water needs: +8/+9% 2025-2055 HadCM3 (B2) Parnaiba (1) -80% 2050s HadCM3 (A2) Palmer et al. (2008) Mimoso catchment (1) Dry scenario: -25/-75%; 2010 2039, 2040 CSMK3 and HadCM3 (A2, B1) Montenegro and Ragab (2010) (2) Wet scenario: +40/+140%; (3) Similar 2069, and 2070 2099 Tapacurá River (1) B1: -4.89%, -14.28%, -20.58% Three 30-year CSMK3 and MPEH5 (A2, B1) Montenegro and Ragab (2012) (1) A2: +25.25%, 39.48% and 21.95% periods Bengue catchment (1) -15% reservoir yield Sensitivity scenario in 2100 selected from TAR and AR4 Krol et al. (2011) GCMs with good skill. + 15% PET, -10% Precip Aquifer in NEB (3) Reduction 2040-2070 HadCM3, ECHAM4 (A2,B2) Hirata and Conicelli (2012) North SA Essequibo Guyana (1) -50% 2050s HadCM3 (A2) Palmer et al. (2008) Magdalena (1) Not significant changes in near future. End 2015 2035 and CMIP3 multi-model ensemble (A1B) Nakaegawa and Vergara (2010) (Colombia) of 21st changes in seasonality. 2075 2099 Sinu (Colombia) (1) -2/-35% 2010-2039 CCSRNIES, CSIROMK2B, CGCM2, Ospina-Norena et al.(2009a; 2009b) HadCM3 (A2) CA Lempa (1) B1: -13%; (1) A2: -24% 2070-2100 CMIP3 (A2, B1) Maurer et al. (2009) Grande de (1) -70% 2050s HadCM3 (A2) Palmer et al. (2008) Matagalpa Mesoamerica (1) Decrease across the region 2070-2100 CMIP3 (A2, A1B, B1) Imbach et al. (2012) (1) Consistent decrease 2050s HadCM3 and CMIP3 (A1b) Arnel and Gosling (2013) (1) Consistent reduction in Northern CA 2050 2099 30 GMCs (A1b) Hidalgo et al. (2013) Panama (1) The  Pacific  basins:  +35/40%   2075 2099 MRI-­ AGCM3.1  (A1B) Fábrega et al. (2013) (1) Bocas del Toro: -50% Subject to Final Copyedit 92 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 27-5: Impacts on agriculture. Country/ Activity Time slice SRES CO2 Changes Source Region Uruguay Annual 2030/2050/2070/2100 A2 +185/-194/-284/-508 ECLAC (SESA) crops 2030/2050 B2 +92/+169 (2010) 2030/2050/2070/2100 A2 +174/-80/-160/-287 (*1) Livestock 2030/2050 B2 +136/+182 2030/2050/2070/2100 A2 +15/+39/+52/+19 Forestry 2030/2050/2070 B2 +6/+13/+18 (Paraguay Cassava 2020/2050/2080 A2 +16/+22/+22 ECLAC (SESA) Wheat 2020/2050/2080 A2 +4/-9/-13 (2010) B2 -1/+1/-5 Maize 2020/2050/2080 A2 +3/+3/+8 B2 +3/+1/+6 A2 Soybean 2020/2050/2080 A2 0/-10/-15 B2 0/-15/-2 Bean 2020/2050/2080 A2 -1/+10/+16 Argentina Maize 2080 A2/B2 N -24/-15 ECLAC (SESA) A2/B2 Y +1/0 (2010) Soybean 2080 A2/B2 N -25/-14 A2/B2 Y +14/+19 Wheat 2080 A2/B2 N -16/-11 A2/B2 Y +3/+3 Soybean 2020/2050/2080 A2 Y +24/+42/+48 Travasso et B2 Y +14/+30/+33 al. (2008) Maize 2020/2050/2080 A2 Y +8/+11/+16 B2 Y +5/+5/+9 Brazil Rice 2CO2/0C Y +60 Walter et al. (SESA) 2CO2/+5C Y +30 (2010) Bean 2050-2080 A2 N Up to -30% Costa et al. 2020-2050-2080 A2+CO2 Y Up to: +30/+30/+45 (2009) 2020-2050-2080 A2+CO2+T Y Up to: +45/+75/+90 (*2) Maize 2050-2080 A2 N Up to -30% 2050-2080 A2+CO2 Y Near to -15% 2020-2050-2080 A2+CO2+T Y Up to: +40/+60/+90 Arabica 0 to +1C +1.5% Zullo et al. coffee +1 to +2C +15.9% (2011) +2 to +3C +28.6% (*3) +3 to +4C -12.9% Brazil Sugarcane 2040 Pessimistic +6% Marin et al. Sao Pablo 2040 Optimistic +2% (2009) Brazil Cassava 2030 N 0 to -10 Lobell et al. Northeast Maize 2030 N 0 to -10 (2008) Rice 2030 N -1 to -10 Wheat 2030 N -1 to -14 Maize -20 to -30 Margulis et Bean -20 to -30 al. (2010) Rice -20 to -30 Cowpea +1.5C -26% Silva et al. bean +3.0C -44% (2010) +5.0C -63% (*3) Central America Maize A2 0/0/-10/-30 ECLAC (CA) Bean 2030/2050/2070/2100 A2 -4/-19/-29/-87 (2010) Rice A2 +3/-3/-14/-63 Rice 2020-2040 N 0 to -10 Lobell et al. Wheat 2020-2040 N -1 to -9 (2008) Panamá Maize 2020-2050-2080 A2 Y -0.5/+2.4/+4.5 Ruane et al. 2020-2050-2080 B1 Y -0.1/-0.8/+1.5 (2011) Andean Region Wheat N -14 to +2 Lobell et al. 2020-2040 Barley N -1 to -8 (2008) Subject to Final Copyedit 93 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Potato N 0 to -5 Maize N 0 to -14 Colombia All main 2050 17GCM-A2 80% of crops Ramirez et crops impacted in more al. (2012) than 60% of current cultivated areas Chile Maize 2050 A1FI Y -5% to -10% Meza and 34.6S/38.5S Wheat 2050 A1FI Y -10% to -20% Silva (2009) Changes are expressed as differences in relative yield (%), except for (*1) and (*3) N: Without considering CO2 biological effects; Y: Considering CO2 biological effects (*1) Gross Value of Production (millions of American dollars) (*2) Huge spatial variability, the values are approximated (*3) Changes in the percentage of areas with low climate risk Subject to Final Copyedit 94 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 27-6: Comparison of consumption of different energetics in Latin America and the world (in thousand tonnes of oil equivalent (ktoe) on a net calorific value basis). LATAM World Energy resource TFC (via TFC (via TFC (non TFC (non electricity Total TFC electricity TFC electricity) electricity) generation) generation) Coal and Peat 9,008 3% 1,398 2% 10,406 3% 831,897 12% 581,248 40% 1,413,145 17% Fossil Oil 189,313 55% 8,685 13% 197,998 48% 3,462,133 52% 73,552 5% 3,535,685 44% Natural Gas 59,44 17% 9,423 14% 68,863 17% 1,265,862 19% 307,956 21% 1,573,818 19% Nuclear Nuclear 0 0% 1,449 2% 1,449 0% 0 0% 193,075 13% 193,075 2% Biofuels and 82,997 24% 2,179 3% 85,176 21% 1,080,039 16% 20,63 1% 1,100,669 14% waste Hydro 0 0% 45,92 66% 45,92 11% 0 0% 238,313 17% 238,313 3% Renewable Geothermal, solar, wind, 408 0% 364 1% 772 0% 18,265 0% 26,592 2% 44,857 1% other renewable TOTAL 341,166 100% 69,418 100% 410,584 100% 6,658,196 100% 1,441,366 100% 8,099,562 100% * TFC: Total final consumption                       Source: IEA, 2012                     Subject to Final Copyedit 95 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 27-7: Cases of government-funded PES schemes in CA and SA. Countries Level Start Name Benefits References Brazil Sub-national 2007 Bolsa Floresta By 2008, 2700 traditional and indigenous Viana (2008) (Amazonas families already benefitted: financial state) compensation and health assistance in exchange for zero deforestation in primary forests. Costa Rica National 1997 FONAFIFO PES is a strong incentive for reforestation Montagnini and fund and, for agroforestry ecosystems alone, Finney (2011) over 7,000 contracts have been set since 2003, and nearly 2 million trees were planted. Ecuador National 2008 Socio-Bosque By 2010, the program already included De Koning et al. more than half a million hectares of natural (2011) ecosystems protected and has over 60,000 beneficiaries. Guatemala National 1997 Programa de By 2009, the program included 4,174 Instituto Nacional Incentivos beneficiaries who planted 94,151 hectares de Estadística Forestales, of forest. In addition, 155,790 hectares of (2011) PINFOR natural forest were under protection with monetary incentives. Table 27-8: Key risks from climate change and the potential for risk reduction through mitigation and adaptation. Subject to Final Copyedit 96 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 27-1: Observed and simulated variations in past and projected future annual average temperature over the Central and South American regions defined in IPCC (2012). Black lines show various estimates from observational measurements. Shading denotes the 5-95 percentile range of climate model simulations driven with "historical" changes in anthropogenic and natural drivers (63 simulations), historical changes in "natural" drivers only (34), the "RCP2.6" emissions scenario (63), and the "RCP8.5" (63). Data are anomalies from the 1986-2006 average of the individual observational data (for the observational time series) or of the corresponding historical all- forcing simulations. Further details are given in Box 21-3. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 97 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 27-2: Projected changes in annual average temperature and precipitation. CMIP5 multi-model mean projections of annual average temperature changes (left panel ) and average percent change in annual mean precipitation (right panel) for 2046-2065 and 2081-2100 under RCP2.6 and 8.5. Solid colors indicate areas with very strong agreement, where the multi-model mean change is greater than twice the baseline variability, and>90% of models agree on sign of change. Colors with white dots indicate areas with strong agreement, where>66% of models show change greater than the baseline variability and>66% of models agree on sign of change. Gray indicates areas with divergent changes, where>66% of models show change greater than the baseline variability, but<66% agree on sign of change. Colors with diagonal lines indicate areas with little or no change, less than the baseline variability in>66% of models. (There may be significant change at shorter timescales such as seasons, months, or days.). Analysis uses model data and methods building from WGI AR5 Figure SPM.8. See also Annex I of WGI AR5. [Boxes 21-3 and CC-RC] Subject to Final Copyedit 98 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 27-3: Area deforested per year for selected countries in CA and SA (2005-2010). Notice three countries listed with a positive change in forest cover (based on data from FAO, 2010). Figure 27-4: Deforestation rates in the Brazilian Amazonia (km/year) based on measurements by the PRODES project (INPE, 2011). Subject to Final Copyedit 99 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 27-5: Evolution of GDP per capita and poverty (income below US$ 2 per day) from 1990-2010: CA and SA (US-Dollars per inhabitant at 2005 prices and percentages) (ECLAC on the basis of CEPALSTAT (2012) and ECLAC (2011b)). Figure 27-6: Current and predicted coastal impacts and coastal dynamics in response to climate change. Coastal impacts - based on trends observed and projections, the figure shows how potential impacts may be distributed in the region. Three cases: a) flooding: since flooding probability increases with increasing sea-level, one may expect a higher probability of flooding in locations showing >40% of change over the last 60 years in the 100- years total sea-level (excluding hurricanes). The figure also identifies urban areas where the highest increase in flooding level has been obtained.; b) beach erosion: it increases with potential sediment transport, thus locations where changes in potential sediment transport have increased over a certain threshold have a higher probability to be eroded; c) sea-ports and reliability of coastal structures: the figure shows locations where, in the case of having a protection structure in place, there is a reduction in the reliability of the structures due to the increase in the design wave height estimates (ECLAC, 2011a). Coastal dynamics - information based on historical time series that have been obtained by a combination of data reanalysis, available instrumental information and satellite information. Advanced statistical techniques have been used for obtaining trends including uncertainties (Izaguirre et al., 2013; Losada et al., 2013). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 100 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 27-7: Summary of observed changes in climate and other environmental factors in representative regions of CA and SA. The boundaries of the regions in the map are conceptual (neither geographic nor political precision). Information and references to changes provided are presented in different sections of the chapter. Subject to Final Copyedit 101 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 27 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 27-8: Observed impacts of climate variations and attribution of causes in CA and SA. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 102 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 28. Polar Regions Coordinating Lead Authors Joan Nymand Larsen (Iceland), Oleg A. Anisimov (Russian Federation) Lead Authors Andrew Constable (Australia), Anne Hollowed (USA), Nancy Maynard (USA), Pal Prestrud (Norway), Terry Prowse (Canada), John Stone (Canada) Contributing Authors Terry Callaghan (UK), Mark Carey (USA), Peter Convey (UK), Andrew Derocher (Canada), Peter T. Fretwell (UK), Bruce C. Forbes (Finland), Solveig Glomsrd (Norway), Dominic Hodgson (UK), Eileen Hofmann (USA), Grete K. Hovelsrud (Norway), Gita L Ljubicic (Canada), Harald Loeng (Norway), Eugene Murphy (UK), Steve Nicol (Australia), Anders Oskal (Norway), James D. Reist (Canada), Phil Trathan (UK), Barbara Weinecke (Australia), Fred Wrona (Canada) Review Editors Maria Ananicheva (Russian Federation), F. Stuart Chapin III (USA) Volunteer Chapter Scientist Vasiliy Kokorev (Russian Federation) Contents Executive Summary 28.1. Introduction 28.2. Observed Changes and Vulnerability under Multiple Stressors 28.2.1. Hydrology and Freshwater Ecosystems 28.2.1.1. Arctic 28.2.1.2. Antarctic 28.2.2. Oceanography and Marine Ecosystems 28.2.2.1. Arctic 28.2.2.2. Antarctica 28.2.3. Terrestrial Ecosystems 28.2.3.1. Arctic 28.2.3.2. Antarctica 28.2.4. Health and Well-Being of Arctic Residents 28.2.4.1. Direct Impacts of a Changing Climate on the Health of Arctic Residents 28.2.4.2. Indirect Impacts of Climate Change on the Health of Arctic Residents 28.2.5. Indigenous Peoples and Traditional Knowledge 28.2.6. Economic Sectors 28.2.6.1. Arctic 28.2.6.2. Antarctica and the Southern Ocean 28.3. Key Projected Impacts and Vulnerabilities 28.3.1. Hydrology and Freshwater Ecosystems 28.3.1.1. Arctic 28.3.1.2. Antarctica 28.3.2. Oceanography and Marine Ecosystems 28.3.2.1. Ocean Acidification in the Arctic and Antarctic 28.3.2.2. Arctic Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 28.3.2.3. Antarctica and the Southern Ocean 28.3.3. Terrestrial Environment and Related Ecosystems 28.3.3.1. Arctic 28.3.3.2. Antarctica 28.3.4. Economic Sectors 28.3.4.1. Fisheries 28.3.4.2. Forestry and Farming 28.3.4.3. Infrastructure, Transportation, and Terrestrial Resources 28.4. Human Adaptation 28.5. Research and Data Gaps References Frequently Asked Questions 28.1: What will be the net socio-economic impacts of change in the polar regions? 28.2: Why are changes in sea ice so important to the polar regions? Executive Summary Additional and stronger scientific evidence has accumulated since the AR4 that reinforces key findings made in the Fourth Assessment Report (AR4). The impacts of climate change, and the adaptations to it, exhibit strong spatial heterogeneity in the Polar Regions because of the high diversity of social systems, bio-physical regions and associated drivers of change (high confidence). [28.2.2] For example, the tree line has moved northward and upward in many, but not all, Arctic areas (high confidence) and significant increases in tall shrubs and grasses have been observed in many places (very high confidence). [28.2.3.2] Some marine species will shift their ranges in response to changing ocean and sea ice conditions in the Polar Regions (medium confidence). The response rate and the spatial extent of the shifts will differ by species based on their vulnerability to change and their life history. [28.2.2; 28.3.2] Loss of sea ice in summer and increased ocean temperature is expected to enhance secondary pelagic production in some but not all regions of the Arctic Ocean with associated changes in the energy pathways within the marine ecosystem (medium confidence). These changes are expected to alter the species composition and carrying capacity with associated impacts on marine fish and shellfish populations (medium confidence). [28.2.2.1] Also, changes in sea ice and the physical environment to the west of the Antarctic Peninsula are altering phytoplankton stocks and productivity, and krill (high confidence). [28.2.2.2] Climate change is impacting terrestrial and freshwater ecosystems in some areas of Antarctica and the Arctic. This is due to ecological effects resulting from reductions in the duration and extent of ice and snow cover and enhanced permafrost thaw (very high confidence), and through changes in the precipitation-evaporation balance (medium confidence). [28.2] The primary concern for polar bears over the foreseeable future is the recent and projected loss of annual ice over continental shelves, decreased ice duration, and decreased ice thickness (high confidence). Of the two subpopulations where data are adequate for assessing abundance effects, it is very likely that the recorded population declines are caused by reductions in sea ice extent. [28.2.2.1.2; 28.3.2.2.2] Rising temperatures, leading to the further thawing of permafrost and changing precipitation patterns have the potential to affect infrastructure and related services in the Arctic (high confidence). Particular concerns Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 are associated with the damage of residential buildings due to thawing permafrost, including, Arctic cities; small, rural settlements; and storage facilities for hazardous materials. [28.2.4; 28.2.4.2; 28.2.5] In addition, there is new scientific evidence that has emerged since the AR4. The physical, biological and socio-economic impacts of climate change in the Arctic have to be seen in the context of often interconnected factors that include not only environmental changes caused by drivers other than climate change but also demography, culture and economic development. Climate change has compounded some of the existing vulnerabilities caused by these other factors (high confidence). [28.2.4; 28.2.5; 28.4] For example, food security for many indigenous and rural residents in the Arctic is being impacted by climate change and in combination with globalization and resource development projected to increase significantly in the future (high confidence). [28.2.4] The rapid rate at which climate is changing in the Polar Regions will impact natural and social systems (high confidence) and may exceed the rate at which some of their components can successfully adapt (low to medium confidence). [28.2.4; 28.4] The decline of Arctic sea-ice in summer is occurring at a rate that exceeds most model projections (high confidence) and evidence of similarly rapid rates of change is emerging in some regions of Antarctica. [IPCC WGI Chapter 14] In the future, trends in Polar Regions of populations of marine mammals, fish and birds will be a complex response to multiple stressors and indirect effects (high confidence). [28.3.2] Already, accelerated rates of change in permafrost thaw, loss of coastal sea ice, sea level rise and increased weather intensity are forcing relocation of some indigenous communities in Alaska (high confidence). [28.2.4.2; 28.2.5; 28.3.4] Shifts in the timing and magnitude of seasonal biomass production could disrupt matched phenologies in the food webs, leading to decreased survival of dependent species (medium confidence). If the timing of primary and secondary production is no longer matched to the timing of spawning or egg release, survival could be impacted with cascading implications to higher trophic levels. This impact would be exacerbated if shifts in timing occur rapidly (medium confidence). [28.2.2; 28.3.2] Climate change will increase the vulnerability of terrestrial ecosystems to invasions by non-indigenous species, the majority likely to arrive through direct human assistance (high confidence). Ocean acidification has the potential to inhibit embryo development and shell formation of some zooplankton and krill in the Polar Regions with potentially far-reaching consequences to food webs in these regions (medium confidence). Embryos of Antarctic krill have been shown to be vulnerable to increased concentrations of CO2 in the water (high confidence). As well, there is increasing evidence that pelagic mollusks (pteropods) are vulnerable to ocean acidification (medium confidence). [28.2.2; 28.3.2] There is increased evidence that climate change will have large effects on Arctic communities, especially where narrowly based economies leave a smaller range of adaptive choices. [28.2.6.1; 28.4] Some commercial activities will become more profitable while others will face decline. Increased economic opportunities are expected with increased navigability in the Arctic Ocean and the expansion of some land- and freshwater-based transportation networks. [28.2.6.1.3; 28.3.4.3] The informal, subsistence-based economy will be impacted (high confidence). There is high confidence that changing sea-ice conditions will result in more difficult access for hunting marine mammals. [28.2.6.1.6] Although Arctic residents have a history of adapting to change, the complex inter-linkages between societal, economic, and political factors and climatic stresses represent unprecedented challenges for northern communities, particularly as the rate of change will be faster than the social systems can adapt (high confidence). [28.4; 28.4.1; 28.2.5] Impacts on the health and well-being of Arctic residents from climate change are significant and projected to increase especially for many indigenous peoples (high confidence). [28.2.4] These are expected to vary among the diverse settlements, which range from small, remote predominantly indigenous communities to large cities and industrial settlements (high confidence), especially those located in highly vulnerable locations along ocean and river shorelines. [28.2.4] Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 28.1. Introduction Several recent climate impact assessments on Polar Regions have been undertaken, including the synthesis report on Snow, Water, Ice and Permafrost in the Arctic (AMAP, 2011), the State of the Arctic Coast 2010 (2011) reports, the Antarctic Climate and the Environment (Turner et al., 2009, 2013), Arctic Resilience Interim Report 2013 (2013), and the findings of the International Polar Year (IPY) (Krupnick et al., 2011). These reports draw a consistent pattern of climate-driven environmental, societal and economic changes in the Polar Regions in recent decades. In this chapter, we use the scientific literature, including these reports, to consolidate the assessment of the impacts of climate change on Polar Regions from 2007, advance new scientific evidence of impacts and identify key gaps in knowledge on current and future impacts. Previous IPCC reports define the Arctic as the area within the Arctic Circle (66oN), and the Antarctic as the continent with surrounding Southern Ocean south of the polar front, which is generally close to 58S (IPCC, 2007). For the purpose of this report we use the conventional IPCC definitions as a basis, while incorporating a degree of flexibility when describing the Polar Regions in relation to particular subjects. [INSERT FIGURE 28-1 HERE Figure 28-1: Location maps of the north and south polar regions. Credit: P. Fretwell, British Antarctic Survey.] Changes in the physical and chemical environments of the Polar Regions are detailed in IPCC WG-1 report. The Arctic has been warming since the 1980s at approximately twice the global rate demonstrating the strongest temperature changes (~ 1 °C per decade) in winter and spring, and smallest in autumn. Sea ice declined at an average rate of 13% per decade; the Arctic Ocean is projected to become nearly ice-free in summer within this century. The duration of snow cover extent and snow depth are decreasing in North-America while increasing in Eurasia. Since late 1970s permafrost temperatures have increased between 0.5 to 2 °C. In the Southern Hemisphere, the strongest rates of atmospheric warming are occurring in the western Antarctic Peninsula (WAP, between 0.2 and 0.3 °C per decade) and the islands of the Scotia Arc, where there have also been increases in oceanic temperatures and large regional decreases in winter sea ice extent and duration. Warming, although less than WAP, has also occurred in the continental margins near to Bellingshausen Sea, Prydz Bay and the Ross Sea, with areas of cooling in between. Land regions have experienced glacial recession and changes in the ice and permafrost habitats in the coastal margins. The Southern Ocean continues to warm, with increased freshening at the surface due to precipitation leading to increased stratification. In both Polar Regions, surface waters will become seasonally corrosive to aragonite within decades, with some regions being affected sooner than others (Box CC-OA, WGI AR5 Chapter 6). Observations and models indicate that the carbon cycle of the Arctic and Southern Oceans will be impacted by climate change and increased CO2. 28.2. Observed Changes and Vulnerability under Multiple Stressors 28.2.1. Hydrology and Freshwater Ecosystems 28.2.1.1. Arctic Arctic rivers and lakes continue to show pronounced changes to their hydrology and ecology. Previously noted increases in Eurasian Arctic river flow (1936-1999; Peterson et al., 2002) could not, for a similar period (1951- 2000), be attributed with certainty to precipitation changes (Milliman et al., 2008) but has been, including more recent extreme increases (2007), to enhanced poleward atmospheric moisture transport (Zhang et al., 2013). By contrast, decreased flow in high-latitude Canadian rivers (1964-2000; average -10%) does match that for precipitation (Déry and Wood, 2005). Recent data (1977-2007) for 19 circumpolar rivers also indicates an area- weighted average increase of +9.8% ( 17.1 to 47.0%; Overeem and Syvitski, 2010) accompanied by shifts in flow timing, with May snowmelt increasing (avg. 66%) but flow in the subsequent month of peak discharge decreasing (~7%). Across the Russian Arctic, dates of spring maximum discharge have also become earlier, particularly in the most recent [1960-2001] period analyzed (average -5d; range for 4 regions +0.2 to -7.1 d), but no consistent trend exists for magnitude (average -1%; range +21 to -24%; Shiklomanov et al., 2007). Earlier timing was most Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 pronounced in eastern, colder continental climates where increases in air temperature have been identified as the dominant control (Tan et al., 2011). Increases have also occurred in winter low flows for many Eurasian and North American rivers (primarily late 20th century; Smith et al., 2007; St. Jacques and Sauchyn, 2009; Walvoord and Striegl, 2007; Ye et al., 2009), the key exceptions being decreases in eastern North America and unchanged flow in small basins of eastern Eurasia (Rennermalm et al., 2010). Most of such studies suggest permafrost thaw (WGI, Ch.4) has increased winter flow, while others suggest increases in net winter precipitation minus evapotranspiration (Landerer et al., 2010; Rawlins et al., 2009a, b). Insufficient precipitation stations preclude deciphering the relative importance of these factors (WGI, Ch. 2.5.1). The surface-water temperatures of large water bodies has warmed (1985-2009; Schneider and Hook, 2010), particularly for mid- and high latitudes of the northern hemisphere with spatial patterns generally matching those for air temperature. Where water bodies warmed more rapidly than air temperature, decreasing ice cover was suggested as enhancing radiative warming. Paleolimnological evidence indicates that highest primary productivity was associated with warm, ice-free summer conditions and the lowest with periods of perennial ice (Melles et al., 2007). Increasing water temperatures affect planktonic and benthic biomass and lead to changes in species composition (Christoffersen et al. 2008; Heino et al. 2009, Jansson et al. 2010). Reduced ice cover with higher air temperatures and evaporation are responsible for the late 20th c early 21st c desiccation of some Arctic ponds (Smol and Douglas, 2007). Changes have occurred in the size and number of permafrost lakes over the last half-century (Hinkel et al., 2007; Marsh et al., 2009), but their patterns and rates of change are not consistent because of differing thawing states, variations in warming and effects of human activities (Hinket et al. 2007; Prowse and Brown, 2010a). Thawing permafrost affects the biogeochemistry of water entering lakes and rivers (Frey and McClelland, 2009; Kokelj et al., 2009) and their ecological structure and function (Lantz and Kokelj, 2008; Mesquita et al., 2010; Thompson et al., 2008), such as enhancing eutrophication by a shift from pelagic to benthic-dominated production (Thompson et al., 2012). The aquatic ecosystem health and biodiversity of northern deltas is dependent on combined changes in the elevation of spring river ice-jam floods and sea level (Lesack and marsh, 2007, 2010). Diminishing ice shelves (last half- century) have also caused a decline in the number of freshwater epishelf lakes that develop behind them (Veillette et al., 2008; Vincent et al., 2009). While such biophysical dependencies have been established, temporal trends in such river-delta and epishelf lake impacts and their linkages to changing climate remain to be quantified precisely. An interplay of freshwater-marine conditions also affect the timing, growth, run size and distribution of several Arctic freshwater and anadromous fish. Key examples include: the timing of marine exit of Yukon River Chinook salmon (Oncorhynchus tshawytscha; 1961-2009) varied with air and sea temperatures and sea ice cover (Mundy and Evenson, 2011); the growth of young-of-year Arctic cisco (Coregonus autumnalis; 1978 2004) varied in response to lagged sea-ice concentration and Mackenzie River discharge, also indicating that decreased sea-ice concentration and increased river discharge enhanced marine primary production leading to more favorable foraging conditions (Von Biela et al. (2011); and factors that influence the water level and freshening of rivers, as well as the strength, duration and directions of prevailing coastal winds, affect survival of anadromous fishes during coastal migration and their subsequent run size (Fechhelm et al., 2007). 28.2.1.2. Antarctic Biota of Antarctic freshwater systems (lakes, ponds, short streams and seasonally wetted areas) are dominated by benthic microbial communities of cyanobacteria and green algae in a simple food web. Mosses occur in some continental lakes with higher plants absent. Planktonic ecosystems are typically depauperate and include small algae, bacteria and colourless flagellates, with few metazoans and no fish (Quesada and Velázquez, 2012). Recent compilations of single-year datasets have reinforced previous conclusions on the changing freshwater habitats in Antarctica (Verleyen et al., 2012). In regions where the climate has warmed the physical impacts on aquatic Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 ecosystems include loss of ice and perennial snow cover, increasing periods of seasonal open water, increased water column temperatures and changes in water column stratification. In some areas a negative water balance has occured due to increased temperature and changes in wind strength driving enhanced evaporation and sublimation and leading to increased salinity in lakes in recent decades (Hodgson et al., 2006a). In other areas, especially glacial forelands, increased temperatures have led to greater volumes of seasonal meltwater in streams and lakes together with increased nutrient fluxes (high confidence). In both cases the balance between precipitation and evaporation can have detectable effects on lake ecosystems (medium confidence) through changes in water body volume and lake chemistry (Lyons et al., 2006; Quesada et al., 2006). Non-dilute lakes with a low lake depth to surface area ratio are most susceptible to inter-annual and inter-decadal variability in the water balance, as measured by changes in specific conductance (high confidence) (Verleyen et al., 2012). Warming in the northwestern Antarctic Peninsula region has resulted in permafrost degradation in the last c. 50 yr impacting surface geomorphology and hydrology (Bockheim et al., 2013) with the potential to increase soil biomass. 28.2.2. Oceanography and Marine Ecosystems 28.2.2.1. Arctic 28.2.2.1.1. Marine plankton, fish, and other invertebrates Working Group I documents the expected physical and chemical changes that will occur in Arctic marine ecosystems (WGI AR5 Chapters 4, 6, and 11). In addition to climate change, naturally occurring interannual, decadal, and multi-decadal variations in climate will influence the Arctic Ocean and its neighboring high latitude seas (WGII Chapter 5). In recent years (2007-2012) ocean conditions in the Bering Sea have been cold (Stabeno et al., 2012a), while the Barents Sea has been warm (Lind and Ingvaldsen, 2012). In this section we build on previous reviews of observed species responses to climate (Wassman et al. 2011) to summarize the current evidence of the impact of physical and chemical changes in marine systems on the phenology, spatial distribution and production of Arctic marine species. For each type of response, the implications for phytoplankton, zooplankton, fish and shellfish are discussed. The implications of these changes on marine ecosystem structure and function will be the result of the synergistic effects of all three types of biological responses. Phenological response The timing of spring phytoplankton blooms is a function of seasonal light, hydrographic conditions, and the timing of sea ice breakup (Wassman, 2011). In addition to the open water phytoplankton bloom, potentially large ice algal blooms can form under the sea ice (Arrigo, 2012). During the period 1997-2009, a trend towards earlier phytoplankton blooms was detected in approximately 11% of the area of the Arctic Ocean (Kahru et al., 2011). This advanced timing of annual phytoplankton blooms coincided with decreased sea ice concentration in early summer. Brown and Arrigo (2013) studied the timing and intensity of spring blooms in the Bering Sea from 1997-2010 and found that in northern regions, sea ice consistently retreated in late spring and was associated with ice-edge blooms, whereas, in the southern regions the timing of sea ice retreat varied, with ice-edge blooms associated with late ice retreat, and open water blooms associated with early ice retreat. Given the short time series and limited studies, there is medium confidence that climate variability and change has altered the timing and the duration of phytoplankton production. The life cycles of calanoid copepods in the Arctic Ocean and Barents Sea are timed to utilize ice algal and phytoplankton blooms (Falk-Petersen et al. 2009; Sreide et al., 2010; Darnis et al. 2012). Based on a synthesis of existing data, Hunt et al. (2011) hypothesized that in the southeastern Bering Sea, ocean conditions and the timing of sea ice retreat influences the species composition of dominant zooplankton, with lipid rich copepods being more abundant in cold years. Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 There is ample evidence that the timing of spawning and hatching of some fish and shellfish is aligned to match larval emergence with seasonal increases availability of prey (Gjosaeter et al., 2009; Vikeb et al., 2010; Bouchard and Fortier, 2011; Drinkwater et al., 2011). These regional phenological adjustments to local conditions occurred over many generations (Ormseth and Norcross, 2009; Geffen et al., 2011; Kristiansen et al., 2011). There is medium to high confidence that climate induced disruptions in this synchrony can result in increased larval or juvenile mortality or changes in the condition factor of fish and shellfish species in the Arctic marine ecosystems. Observed spatial shifts Spatial heterogeneity in primary production has been observed (Lee et al., 2010; Grebmeier, 2012). Simulation modeling studies show that spatial differences in the abundance of four species of copepod can be explained by regional differences in the duration of the growing season and temperature (Ji et al., 2012). Retrospective studies based on surveys from 1952-2005 in the Barents Sea revealed that changes in the species composition, abundance and distribution of euphausiids were related to climate-related changes in oceanographic conditions (Zhukova et al., 2009). Retrospective analysis of observed shifts in the spatial distribution of fish and shellfish species along latitudinal and depth gradients showed observed spatial shifts were consistent with expected responses of species to climate change (Simpson et al., 2011; Poloczanska et al. 2013; Box CC-MB, Chapter 30). Retrospective studies from the Bering Sea, Barents Sea, and the northeast Atlantic Ocean and Icelandic waters, showed that fish shift their spatial distribution in response to climate variability (i.e. interannual, decadal or multi-decadal changes in ocean temperature; Mueter and Litzow, 2008; Sundby and Nakken, 2008; Hátún et al., 2009; Valdimarsson et al., 2012; Kotwicki and Lauth, 2013). There are limits to the movement potential of some species. Vulnerability assessments indicate that the movement of some sub-arctic fish and shellfish species into the Arctic Ocean may be impeded by the presence of water temperatures on the shelves that fall below their thermal tolerances (Hunt et al., 2013; Hollowed et al., 2013). Coupled bio-physical models have reproduced the observed spatial dynamics of some the species in the Bering and Barents Seas, and are being used to explain the role of climate variability and change on the distribution and abundance of some species (Huse and Ellingsen, 2008; Parada et al., 2010). In summary, there is medium to high confidence based on observations and modeling that some fish and shellfish have shifted their distribution in response to climate impacts on the spatial distribution and volume of suitable habitat. Observed variations in production Seasonal patterns in light, sea ice cover, freshwater input, stratification and nutrient exchange act in concert to produce temporal cycles of ice algal and phytoplankton production in Arctic marine ecosystems (Wassmann, 2011; Perrette et al., 2011; Tremblay et al., 2012). Satellite observations and model estimates for the period 1988 2007 showed that phytoplankton productivity increased in the Arctic Ocean in response to a downward trend in the extent of summer sea ice (Zhang et al., 2010). Satellite data provided evidence of a 20% increase in annual net primary production in the Arctic Ocean between 1998 and 2009 in response to extended ice free periods (Arrigo and van Dijken, 2011). Regional trends in primary production will differ in response to the amount of open water area in summer (Arrigo and van Dijken, 2011). Other studies showed gross primary production increased with increasing air temperature in the Arctic Basin and Eurasian shelves (Slagstad et al., 2011). A recent 5 year study (2004-2008) in the Canadian Basin showed that smaller phytoplankton densities were higher than larger phytoplankton densities in years when sea surface temperatures were warmer, the water column was more stratified, and nutrients were more depleted during the Arctic summer (Li et al., 2009; Morán et al., 2010). Additional observations will help to resolve observed differences between in-situ and satellite derived estimates of primary production (Matrai et al., 2013). In conclusion, based on recent observations and modeling, there is medium to high confidence that primary production has increased in the Arctic Ocean in response to changes in climate and its impact on the duration and areal extent of ice free periods in summer. Regional differences in zooplankton production have been observed. During a period of ocean warming (1984- 2010), Dalpadado et al. (2012) observed an increase in the biomass of lipid rich euphausiids in the Barents Sea and Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 relatively stable levels of biomass and production of Calanus finmarchicus. In the Bering Sea, observations over the most recent decade in the southeast Bering Sea showed C. marshallae were more abundant in cold years than warm years (Coyle et al., 2011). There is strong evidence that climate variability impacts the year-class strength of Arctic marine fish and shellfish through its influence on: predation risk; the quality, quantity and availability of prey; and reproductive success (Mueter et al., 2007; Bakun 2010; Drinkwater et al., 2010). Regional differences in the species responses to climate change will be a function of the exposure of the species to changing environmental conditions, the sensitivity of the species to these changes (Beaugrand and Kirby, 2010) and the abilities of species to adapt to changing conditions (Pörtner and Peck, 2010; Donelson et al., 2011). There is high confidence that shifts in ocean conditions have impacted the abundance of fish and shellfish in Arctic regions. Observed trends in the abundance of commercial fish and shellfish may also be influenced by historical patterns of exploitation (Vert-pre et al., 2013). 28.2.2.1.2. Marine mammals, polar bear, and sea birds Studies on responses of Arctic and subarctic marine mammals to climate change are limited and vary according to insight into their habitat requirements and trophic relationships (Laidre et al., 2008). Many Arctic and subarctic marine mammals are highly specialized, have long-life spans, and are poorly adapted to rapid environmental change (Moore and Huntington, 2008), and changes may be delayed until significant sea ice loss has occurred (Freitas et al., 2008; Laidre et al., 2008). Climate change effects on Arctic and subarctic marine mammal species will vary by life history, distribution, and habitat specificity (high confidence). Climate change will improve conditions for a few species, have minor negative effects for others, and some will suffer major negative effects (Ragen et al., 2008; Laidre et al., 2008). Climate change resilience will vary and some ice-obligate species should survive in regions with sufficient ice and some may adapt to ice-free conditions (Moore and Huntington, 2008). Less ice-dependent species may be more adaptable but an increase in seasonally migrant species could increase competition (Moore and Huntington, 2008). Climate change vulnerability was associated with feeding specialization, ice dependence, and ice reliance for prey access and predator avoidance (Laidre et al., 2008). There is medium agreement on which species life histories are most vulnerable. Hooded seals (Cystophora cristata) and narwhal (Monodon monoceros) were identified as most at risk and ringed seals (Pusa hispida) and bearded seals (Erignathus barbatus) as least sensitive (Laidre et al., 2008). Kovacs et al. (2010) shared concern for hooded seals and narwhal but had concerns for ringed seals and bearded seals. Narwhal may have limited ability to respond to habitat alteration (Williams et al. 2011). Species that spend only part of the year in the Arctic (e.g., gray whale (Eschrichtius robustus), killer whale (Orcinus orca)) may benefit from reduced ice (Moore, 2008; Laidre et al., 2008; Higdon and Ferguson, 2009; Matthews et al., 2011; Ferguson et al., 2012). Killer whale expansion into the Arctic could cause a trophic cascade (Higdon and Ferguson, 2009) although there is limited evidence at this time. There is limited evidence although medium agreement that generalists and pelagic feeding species may benefit from increased marine productivity from reduced ice while benthic feeding species near continental shelf habitats may do poorly (Bluhm and Gradinger, 2008). There is limited evidence but high agreement that dietary or habitat specialists will do poorly with reduced ice. Reduction of summer/autumn ice was the primary extrinsic factor affecting Pacific walrus (Odobenus rosmarus) with predictions of distribution changes, reduced calf recruitment, and longer-term predictions of high extinction probability (Cooper et al., 2006; MacCracken, 2012). Summer ice retreat may make migration to such habitats energetically unprofitable for ringed seals (Freitas et al., 2008). Ice loss threatens Baltic ringed seals (Kovacs and Lydersen, 2008). In Hudson Bay, earlier spring break-up and changes in snow cover over lairs have reduced ringed seal recruitment (Ferguson et al., 2005). Changes in snowfall over the 21st century were projected to reduce ringed seal habitat for lairs by 70% (Hezel et al., 2012). Similarly, harp seal (Pagophilus groenlandicus) breeding habitat was affected by changing ice conditions that could reduce pup survival (Bajzak et al., 2011). While there is limited evidence, there are concerns that climate change may cause indirect effects on Arctic marine mammals health (e.g., pathogen transmission, food web changes, toxic chemical exposure, shipping, and development) (Burek et al., 2008). Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Empirical studies provide direct insight into the mechanisms of climate change impact on polar bears (Ursus maritimus) but modelling allows predictive capacity (Hunter et al., 2010; Amstrup et al., 2010; Durner et al., 2011; Castro de la Guardia et al., 2013). Polar bears are highly specialized and use annual ice over the continental shelves as their preferred habitat (Durner et al., 2009; Miller et al., 2012). The recent and projected loss of annual ice over continental shelves, decreased ice duration, decreased ice thickness, and habitat fragmentation is causing reduced food intake, increased energy expenditure, and increased fasting in polar bears (high confidence) (Stirling and Parkinson, 2006; Regehr et al., 2007; Durner et al., 2009; Amstrup et al., 2010; Hunter et al., 2010; Derocher et al., 2011; Rode et al., 2012; Sahanatien and Derocher, 2012; Castro de la Guardia et al., 2013). Subpopulation response varies geographically. Only 2 of the 19 subpopulations, Western Hudson Bay (Regehr et al., 2007) and the Southern Beaufort Sea (Regehr et al., 2010; Rode et al., 2010a) have data series adequate for clear identification of abundance effects related to climate change. Many other subpopulations show characteristics associated with decline but some remain stable. Declining ice is causing lower body condition, reduced individual growth rates, lower fasting endurance, lower reproductive rates, and lower survival (high confidence) (Regehr et al., 2007; Regehr et al., 2010; Rode et al., 2010a; Molnar et al., 2011; Rode et al., 2012). Condition is a precursor to demographic change (very high confidence) (Regehr et al., 2010; Hunter et al., 2010; Rode et al., 2010a; Robinson et al., 2011). The decline in the subpopulation in Western Hudson Bay by 21% between 1987 and 2004 was related to climate change (medium confidence) (Regehr et al., 2007). Replacement of multiyear ice by annual ice could increase polar bear habitat (low confidence) (Derocher et al., 2004). Increasing the distance to multiyear ice and terrestrial refugia at maximal melt may result in drowning, cub mortality, and increased energetic costs (Monnett and Gleason, 2006; Durner et al., 2011; Pagano et al., 2012). There is robust evidence of changes in sea ice conditions changing polar bear distribution including den areas (high confidence) (Fischbach et al., 2007; Schliebe et al., 2008; Gleason and Rode, 2009; Towns et al., 2010; Derocher et al., 2011). The number of human-bear interactions are projected to increase with warming (high confidence) (Stirling and Parkinson, 2006; Towns et al., 2009). Use of terrestrial resources by polar bears was suggested as adaptive (Dyck et al., 2007; Dyck and Romberg, 2007; Armstrong et al., 2008; Dyck et al., 2008; Dyck and Kebreab, 2009; Rockwell and Gormezano, 2009; Smith et al., 2010). Polar bears cannot adapt to terrestrial foods (Stirling et al., 2008b; Amstrup et al., 2009; Slater et al., 2010; Rode et al., 2010b), and will most likely not be able to adapt to climate change and reduced sea ice extent (very high confidence). Changing ice conditions are linked to cannibalism (Amstrup et al., 2006), altered feeding (Cherry et al., 2009), unusual hunting behaviour (Stirling et al., 2008a), and diet change (Iverson et al., 2006; Thiemann et al., 2008) (medium confidence). Upwelling or subsurface convergence areas found in frontal zones and eddies, and the marginal ice zone, are associated with high marine productivity important to Arctic seabirds (e.g. (Irons et al., 2008). Long-term or permanent shifts in convergence areas and the marginal ice-edge zone induced by climate change may cause mismatch between the timing of breeding and the peak in food availability and, thus, potentially have strong negative impacts on seabird populations (Gaston et al., 2005; Moline et al., 2008; Gaston et al., 2009; Gremillet and Boulinier, 2009) (medium confidence). The contrasting results from the relatively few studies of impacts of climate change on Arctic seabirds, demonstrate that future impacts will be highly variable between species and between populations of the same species (medium confidence). Retreating sea ice and increasing SSTs have favored some species and disadvantaged others (Gaston et al. 2005; Byrd et al. 2008; Irons et al. 2008; Karnovsky et al. 2010; Fredriksen et al. 2013). Some species of sea birds respond to a wide range of sea surface temperatures via plasticity of their foraging behaviour, allowing them to maintain their fitness levels (Grémillet et al. 2012). Phenological changes and changes in productivity of some breeding colonies have been observed (Byrd et al. 2008; Gaston and Woo 2008; Moe et al. 2009). Negative trends in population size, observed over the last few decades for several species of widespread Arctic seabirds, may be Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 related to over-harvesting and pollution as well as climate change effects (Gaston, 2011). For those species whose distribution is limited by sea ice and cold water, polar warming could be beneficial (Mehlum 2012). A major ecosystem shift in the Northern Bering Sea starting in the mid 1990s caused by increased temperatures and reduced sea ice cover had a negative impact on benthic prey for diving birds and these populations have declined in the area (Grebmeier et al., 2006). More recently, the Bering Sea has turned colder again. 28.2.2.2. Antarctica Productivity and food web dynamics in the Southern Ocean are dominated by the extreme seasonal fluctuations of irradiance and the dynamics of sea ice, along with temperature, carbonate chemistry and vertical mixing (Massom and Stammerjohn, 2010; Boyd et al., 2012; Murphy et al., 2012a). Moreover, there is large-scale regional variability in habitats (Grant et al., 2006) and their responses to climate change (WGI). Antarctic krill, Euphausia superba (hereafter, krill ) is the dominant consumer, eating diatoms, and, in turn, is the main prey of fish, squid, marine mammals and seabirds. Krill is dominant from the Bellingshausen Sea east through to the Weddell Sea and the Atlantic sector of the Southern Ocean (Rogers et al., 2012). In the East Indian and southwest Pacific sectors of the Southern Ocean, the krill-dominated system lies to the south of the Southern Boundary of the Antarctic Circumpolar Current (Nicol et al., 2000a,b) while to the north copepods and myctophid fish are most important (Rogers et al., 2012). Further west, where the Weddell Sea exerts an influence, krill are found as far north as the Subantarctic Circumpolar Current Front (Jarvis et al., 2010). Where sea ice dominates for most of the year, ice-obligate species (e.g. Euphausia crystallorophias and Peluragramma antarcticum) are most important (Smith et al., 2007). Few studies were available in AR4 to document and validate the changes in these systems resulting from climate change. Those studies reported increasing abundance of benthic sponges and their predators, declining populations of krill, Adélie and emperor penguins, and Weddell seals, and a possible increase in salps, noting some regional differences in these trends. The importance of climate processes in generating these changes could not be distinguished from the indirect consequences of the recovery of whale and seal populations from past over- exploitation (Trathan and Reid, 2009; Murphy et al., 2012a,b). 28.2.2.2.1. Marine plankton, krill, fish, and other invertebrates Distributions of phytoplankton and zooplankton have moved south with the frontal systems (Hinz et al., 2012; Mackey et al., 2012), including range expansion into the Southern Ocean from the north by the coccolithophorid, Emiliania huxleyi (Cubillos et al., 2007), and the red-tide dinoflagellate Noctiluca scintillans (McLeod et al., 2012) (medium confidence). There is insufficient evidence to determine whether other range shifts are occurring. Collapsing ice shelves are altering the dynamics of benthic assemblages by exposing areas previously covered by ice shelves, allowing increased primary production and establishment of new assemblages (e.g. collapse of the Larson A/B ice shelves) (Peck et al., 2009; Gutt et al., 2011) (medium confidence). More icebergs are grounding, causing changes in local oceanography and declining productivity that consequently affects productivity of benthic assemblages (Thrush and Cummings 2011) (low confidence). Iceberg scour on shallow banks is also increasing, disrupting resident benthic assemblages (Barnes and Souster 2011; Gutt et al., 2011) (medium confidence). Primary production is changing regionally in response to changes in sea ice, glacial melt and oceanographic features (Arrigo et al., 2008; Boyd et al., 2012) (medium confidence). Off the west Antarctic Peninsula, phytoplankton stocks and productivity have decreased north of 63°S, but increased south of 63°S (Montes-Hugo et al., 2009) (WGII Chapter 6) (high confidence). This study (based on time-series of satellite-derived and measured chlorophyll concentrations) also indicated a change from diatom-dominated assemblages to ones dominated by smaller phytoplankton (Montes-Hugo et al., 2009). The reduced productivity in the north may be tempered by increased inputs of iron through changes to ocean processes in the region (Dinniman et al., 2012) (low confidence). Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Since the 1980s, Antarctic krill densities have declined in the Scotia Sea (Atkinson et al., 2004), in parallel with regional declines in the extent and duration of winter sea ice (Flores et al., 2012). Uncertainty remains over changes in the krill population because this decline was observed using net samples and is not reflected in acoustic abundance time series (Nicol and Brierley 2010); the observed changes in krill density may have been partly a result of changes in distribution (Murphy et al. 2007). Nevertheless, given its dependence on sea ice (Nicol et al., 2008) the krill population may already have changed and will be subject to further alterations (high confidence). The response of krill populations is probably a complex response to multiple stressors. Decreases in recruitment of post-larval krill across the Scotia Sea have been linked to declines in sea-ice extent in the Antarctic Peninsula region (Wiedenmann et al., 2009) (medium confidence) but these declines may have been offset by increased growth arising from increased water temperature in that area (Wiedenmann et al., 2008). However, near South Georgia krill productivity may have declined as a result of the increased metabolic costs of increasing temperatures (Hill et al., 2013) (low confidence). The combined effects of changing sea ice, temperature and food have not been investigated. 28.2.2.2.2. Marine mammals and sea birds In general, many Southern Ocean seals and seabirds exhibit strong relationships to a variety of climate indices, and many of these relationships are negative to warmer conditions (Trathan et al., 2007; Barbraud et al., 2012; Forcada et al., 2012) (low confidence). Regional variations in climate change impacts on habitats and food will result in a mix of direct and indirect effects on these species. For example, Adélie penguin colonies are declining in recent decades throughout the Antarctic Peninsula while the reduction in chinstrap penguins is more regional (Lynch et al., 2012) and related to reductions in krill availability (Lima and Estay, 2013). In contrast gentoo penguins are increasing in that region and expanding south (Lynch et al., 2012) (high confidence). This may be explained by the reduced sea ice habitats and krill availability in the north resulting in a southward shift of krill predators, particularly those dependent on sea ice (Forcada et al., 2012) and the replacement of these predators in the north by species that do not depend on sea ice, such as gentoo penguins and elephant seals (Costa et al., 2010; Trivelpiece et al., 2011; Ducklow et al., 2012; Murphy et al., 2013) (low confidence). A contrasting situation is in the Ross Sea where Adélie penguin populations have increased (Smith et al., 2012). The mechanisms driving these changes are currently under review and may be more than simply sea ice (Lynch et al., 2012; Melbourne-Thomas et al., 2013). For example, too much or too little sea ice may have negative effects on the demography of Adélie and emperor penguins (see Barbraud et al., 2012 for review). Also, increased snow precipitation which accumulates in breeding colonies can decrease survival of chicks of Adélie penguins when accompanied by reduced food supply (Chapman et al., 2011). Changes elsewhere are less well known. Some emperor penguin colonies have decreased in recent decades (Barbraud et al., 2008; Jenouvrier et al., 2009) (low confidence) and one breeding site has been recorded as having been vacated (Trathan et al., 2011). However, there is insufficient evidence to make a global assessment of their current trend. In the subantarctic of the Indian sector, reductions in seal and seabird populations may indicate a region-wide shift to a system with lower productivity (Weimerskirch et al., 2003; Jenouvrier et al., 2005a, b) (low confidence) but commercial fishing activities may also play a role. Where frontal systems are shifting south, productive foraging areas also move to higher latitudes. In the Indian sector, this is thought to be causing declines in king penguin colonies on subantarctic islands (Péron et al., 2010) (low confidence), while the shift in wind patterns may be causing changes to the demography of albatross (Weimerskirch et al., 2012) (low confidence). As identified in the last assessment, some species populations may suffer as a result of fisheries while others are recoverying from past over-exploitation, either of which may confound interpretation of the response of these species and their food webs to climate change. The recovery of Antarctic fur seals on some subantarctic islands has been well documented and their populations may now be competing with krill-eating macaroni penguins (Trathan et al., 2012). More recently, there has been confirmation that populations of some Antarctic whales are recovering, such as humpbacks (Nicol et al., 2008; Zerbini et al., 2010), suggesting that food is currently not limiting. In contrast, a number of albatross and petrel populations are declining as a result of incidental mortality in longline fisheries in southern and temperate waters where these birds forage (Croxall et al., 2012). Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 28.2.3. Terrestrial Ecosystems 28.2.3.1. Arctic Arctic terrestrial ecosystems have undergone dramatic changes throughout the late Pleistocene and Holocene (last 130 000 years) mainly driven by natural climate change. Significant altitudinal and latitudinal advances and retreats in tree line have been common, animal species have gone extinct, and animal populations have fluctuated significantly throughout this period e.g. (Lorenzen et al., 2011; Salonen et al., 2011; Mamet and Kershaw, 2012). 28.2.3.1.1. Phenology Phenological responses attributable to warming are apparent in most Arctic terrestrial ecosystems (medium confidence). They vary from earlier onset and later end of season in western Arctic Russia (Zeng et al., 2013), to little overall trend in plant phenology in the Swedish sub Arctic (Callaghan et al., 2010), to dramatic earlier onset of phenophases in Greenland (Hye et al., 2007; Post et al., 2009a; Callaghan et al., 2011a) (Figure 28-2). [INSERT FIGURE 28-2 HERE Figure 28-2: Temporal change in onset of flowering (plants), median date of emergence (arthropods) and clutch initiation dates (birds) in high-Arctic Greenland. Red dots are statistically significant, blue dots are not. Source: Hye et al., 2007.] 28.2.3.1.2. Vegetation The latest assessment of changes in NDVI (Normalized Difference Vegetation Index, a proxy for plant productivity) from satellite observations between 1982 and 2012 shows that about a third of the Pan-Arctic has substantially greened, <4% browned and > 57% did not change significantly (Xu et al., 2013) (Figure 28-3). The greatest increases reported in recent years were in the North American high Arctic, along the Beaufort Sea and the east European Arctic (Zhang et al., 2008; Pouliot et al., 2009; Bhatt et al., 2010; Forbes et al., 2010; Walker et al., 2011; Epstein et al., 2012; Macias-Fauria et al., 2012; Xu et al., 2013). [INSERT FIGURE 28-3 HERE Figure 28-3: Significant changes (p< 0.01) in photosynthetically active period NDVI between 1982 and 2012. Source: Xu et al., 2013.] The positive trends in NDVI are associated with increases in the summer warmth index (sum of the monthly-mean temperatures above freezing expressed as oC per month) that have increased on average by 5oC per month for the Arctic as a whole (Xu et al., 2013). However, the even greater 10 to 12oC per month increase for the land adjacent to the Chukchi and Bering Seas (Figure 28-3) was associated with decreases in NDVI. On the Yamal Peninsula in Russia the pattern of NDVI is partly due to surface disturbance, such as landslide activity (Walker et al., 2009). Small rodent cycles reduce NDVI in sub Arctic Sweden, by decreasing biomass and changing plant species composition (Olofsson et al., 2012). The changing NDVI signal should therefore generally be interpreted with care. In common with treeline trees and herbs, the abundance and biomass of deciduous shrubs and graminoids (grasses and grass-like plants) have increased substantially in certain parts of the Arctic tundra in recent years, but remained stable or decreased in others (very high confidence). Attribution for the increases and decreases in deciduous shrubs and graminoids is heterogeneous with drivers varying among different regions (very likely), including Arctic warming, differences in herbivory, industrial development, legacies from past land use, and changes in moisture (Post and Pedersen 2008; Olofsson et al., 2009; Forbes et al., 2009 and 2010; Kitti et al., 2009; Kumpula et al., 2011 and 2012; Myers-Smith et al., 2011; Elmendorf et al., 2012b;Callaghan et al., 2011b and 2013: Gamon et al., 2013). Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Shrubs have generally expanded their ranges and/or growth over the last 20 years (Danby and Hik, 2007; Hudson and Henry, 2009; Forbes et al., 2010; Hallinger et al., 2010; Rundqvist et al., 2011; Hedenas et al., 2011; Hill and Henry, 2011; Myers-Smith et al., 2011a,b; Callaghan et al., 2011b; Elmendorf et al., 2012a,b; Macias-Fauria et al., 2012), and have varied from dramatic, i.e. 200% area increase in study plots (Rundqvist et al., 2011) in sub arctic Sweden to early invasion of a fell field community on west Greenland by low shrubs (Callaghan et al., 2011a). A synthesis (61 sites: (Elmendorf et al., 2012a) of experimental warming studies of up to 20 years duration in tundra sites worldwide, showed, overall, increased growth of deciduous shrubs and graminoids, decreased cover of mosses and lichens, and decreased species diversity and evenness. Elmendorf et al., (2012a) point out that the groups that increased most in abundance under simulated warming were graminoids in cold regions and primarily shrubs in warm regions of the tundra. However, strong heterogeneity in responses to the experimental warming suggested that other factors could moderate the effects of climate warming significantly like herbivory, differences in soil nutrients and pH, precipitation, winter temperatures and snow cover, and species composition and density. Snow bed habitats have decreased in sub arctic Sweden (Björk and Molau, 2007; Hedenas et al., 2011). In other plant communities, changes have been less dramatic, ranging from small increases in species richness in the south west Yukon of the Canadian sub Arctic (Danby et al., 2011), through subtle changes in plant community composition in west and southeast Greenland (Daniëls and De Molenaar, 2011; Callaghan et al., 2011a) to 70 year stability of a plant community on Svalbard (Prach et al., 2010). The responses to Arctic warming of lichen and bryophyte (mosses) diversity have been heterogenous, varying from consistent negative effects to significant increases in recent years (Hudson and Henry, 2009; Tmmervik et al., 2009; Tmmervik et al., 2012). Forbes and Kumpula (2009) recorded long-term and widespread lichen degradation in northern Finland attributed more to trampling of dry lichens by reindeer in summer than winter consumption as forage. Palaeorecords of vegetation change indicate that the northern tree line should extend upwards and northwards during current climate warming (Callaghan et al., 2005) because tree line is related to summer warmth (e.g. Harsch et al., 2009). Although the tree line has moved northwards and upwards in many Arctic areas, it has not shown a general circumpolar expansion in recent decades (high confidence). Model projections that suggest a displacement of between 11 and 50% of tundra by forest by 2100 (see references in Callaghan et al., 2005) and shifts upslope by 2 to 6 m per year (Moen et al., 2004) and northwards by 7.4 20 km per year (Kaplan and New, 2006) might be overestimating rate of tree line advance by a factor of up to 2000 (Van Bogaert et al., 2011). The fastest upslope shifts of tree lines recorded during 20th century warming are 1 to 2 m per year (Shiyatov et al., 2007; Kullman and Öberg, 2009) whereas the fastest so-far recorded northward-migrating tree line replaces tundra by taiga at a rate of 3 10 m per year (Kharuk et al., 2006). In some areas, the location of the tree line has not changed or has changed very slowly (Payette, 2007; MacDonald et al., 2008a). A global study by Harsch et al. (2009) showed that only 52% of 166 global tree line sites studied had advanced over the past 100 years. In many cases the tree line has even retreated (Cherosov et al., 2010). At the small scale, the tree line has shown increase, decrease and stability in neighboring locations (Van Bogaert et al., 2011; Lloyd et al., 2011). Evidence for densification of the forest at the sub Arctic tree line is robust and consistent within Fennoscandia (Tmmervik et al., 2009; Rundqvist et al., 2011; Hedenas et al., 2011) and Canada (Danby and Hik, 2007). Dendroecological studies indicate enhanced conifer recruitment during the twentieth century in the northern Siberian taiga (Briffa et al., 2008). Some of the changes are dramatic, such as an increase in area of mountain birch in study plots in northern Sweden by 600% between 1977/8 and 2009/10 (Rundqvist et al., 2011) and a doubling of tree biomass in Finnmarksvidda in northern Norway since 1957 (Tmmervik et al., 2009). However, model projections of displacement of deciduous forest by evergreen forest (Wolf et al., 2008; Wramneby et al., 2010) have not so far been validated. Where the mountain birch tree line has increased in elevation and shrub (e.g. willow, dwarf birch) abundance has increased, the response can be an interaction between climate warming, herbivory pressure and earlier land use Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (Olofsson et al., 2009; Hofgaard et al., 2010; Van Bogaert et al., 2011). In Fennoscandia and Greenland, heavy grazing by large herbivores may significantly check deciduous low erect shrub (e.g. dwarf shrub and willow) growth (Post et al., 2008; Kitti et al., 2009; Olofsson et al., 2009). Less moisture from snow and more rain now favors broadleaf trees over conifers and mosses in some areas (Juday, 2009) while moisture deficits are reducing the growth of some northern forests (Goetz et al., 2005; Verbyla, 2008; Yarie, 2008) and making them more susceptible to insect pest outbreaks (see references in (Callaghan et al., 2011c). Death of trees through drought stress or insect pest activity will increase the probability of fire that will have positive feedbacks (increase warming) on the climate (Mack et al., 2011). 28.2.3.1.3. Changes in animal populations The documented collapse or dampening of population cycles of voles and lemmings over the last 20-30 years in parts of Fennoscandia and Greenland (Schmidt et al., 2012), can be attributed with high confidence to climate change (Ims et al., 2007; Gilg et al., 2009; Ims et al., 2011; Kausrud et al., 2009). A shortening of the snow season and more thaw and/or rain events during the winter will influence on the subnivean space which provide thermal insulation, access to food, and protection from predators (Berg et al., 2008; Johansson et al., 2011; Kausrud et al., 2009). However, the causes of the changes in the lemming and vole cycles are still being debated as factors other than climate change may also be of importance (Brommer et al., 2010; Krebs, 2011). Climate-mediated range expansion both in altitude and latitude of insect pests, and increased survival due to higher winter temperatures, has been documented for bark beetles in North America (Robertson et al., 2009) and for geometrid moths in Fennoscandia (Jepsen et al., 2008; Callaghan et al., 2010; Jepsen et al., 2011), causing more extensive forest damage than before. Outbreaks of insect pests like geometrid moths can even reduce the strengths of CO2 sinks in some areas (Heliasz et al., 2011). The decline in wild reindeer and caribou (both Rangifer tarandus) populations in some regions of about 30 percent over the last 10-15 years has been linked both to climate warming and anthropogenic landscape changes (Post et al., 2009a; Vors and Boyce 2009; Russell and Gunn 2010). Even though most of the Arctic has warmed, the decline in the populations has not been uniform. Some of the North American large, wild herds have for example declined by 75-90 percent, while other wild herds and semi-domestic herds in Fennoscandia and Russia have been stable or even increased (Gunn et al., 2009; Vors and Boyce, 2009; Joly et al., 2011; Forbes et al., 2009; Forbes 2010; Kumpula et al., 2012). The expected and partially observed increased primary productivity of Arctic tundra may potentially increase the supply of food for Arctic ungulates. However, the overall quality of forage may decline during warming, for example if the nitrogen content of key fodder species for ungulates were to drop during warming (Turunen et al., 2009; Heggberget et al., 2010), while lichen biomass, an important winter fodder for reindeer, is decreasing over parts of the Arctic region. Herbivory also changes the vegetation itself in concert with the warming, further complicating the prediction of vegetation changes and their impacts on ungulate populations (van Der Wal et al., 2007; Turunen et al., 2009). More frequent rain-on-snow icing events and thicker snow-packs caused by warmer winters and increased precipitation may restrict access to vegetation and may have profound negative influences on the population dynamics of Arctic ungulates (Berg et al., 2008; Forchhammer et al., 2008; Miller and Barry, 2009; Stien et al., 2010; Hansen et al., 2011; Stien et al., 2012). Such events have caused heavy mortality in some semi-domestic reindeer herds and musk oxen in recent years (Grenfell and Putkonen, 2008; Forbes, 2009; Bartsch et al., 2010), and have also been shown to synchronize the dynamics of a resident vertebrate community (small mammals, reindeer and Arctic fox) in Svalbard (Hansen et al., 2013). In contrast, Tyler et al. (2008) and Tyler (2010) suggested that generally warmer winters enhance the abundance of reindeer populations. Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 It has been suggested that warming-induced trophic mismatches between forage availability and quality and timing of calving have a role in the decline of circumpolar reindeer and caribou populations (Post and Forchhammer, 2008; Post et al., 2009a,b), although such trophic mismatch has been disputed (Griffith et al., 2010). Adjustment via phenotypic plasticity instead of adaptation by natural selection is expected to dominate vertebrate responses to rapid Arctic climate change, and many such adjustments have already been documented (Gilg et al., 2012). 28.2.3.1.4. Long-term trends and event-driven changes Long-term climate change impacts on vegetation and animal populations are accelerated when tipping points are triggered by events such as extreme weather, fire, insect pest and disease outbreaks. The impacts of winter thaw events on ecosystems are now well-documented e.g. (Bokhorst et al., 2011) but studies of the severe impacts of tundra fires on vegetation and biospheric feedbacks are recent (Mack et al., 2011). Results from experimental winter thaws were validated by a natural event in northern Norway and Sweden in 2007 that reduced NDVI by almost 30% over at least 1400 km2 (Bokhorst et al., 2009). Studies on relationships between climate change and plant disease are rare but Olofsson et al., (2011) showed that increased snow accumulation led to a higher incidence of fungal growth on sub Arctic vegetation. 28.2.3.2. Antarctica Antarctic terrestrial ecosystems occur in 15 biologically-distinct areas (Terauds et al., 2012) with those in the maritime and sub-Antarctic islands experiencing the warmest temperatures, reduced extreme seasonality and greatest biodiversity (Convey, 2006). In the cooler conditions on the continent, species must be capable of exploiting the short periods where temperature and moisture availability are above physiological and biochemical thresholds. In many areas there is no visible vegetation, with life being limited, at the extreme, to endolithic (within rock) communities of algae, cyanobacteria, fungi, bacteria and lichens (Convey, 2006). Few robust studies are available of biological responses to observed climatic changes in natural Antarctic terrestrial ecosystems. The rapid population expansion and local-scale colonisation by two native flowering plants (Deschampsia antarctica and Colobanthus quitensis) in maritime Antarctica (Parnikoza et al., 2009) remains the only published repeat long-term monitoring study of any terrestrial vegetation or location in Antarctica. Radiocarbon dating of moss peat deposits has shown that growth rates and microbial productivity have risen rapidly on the Antarctic Peninsula since the 1960s, consistent with temperature changes, and are unprecedented in the last 150 years (Royles et al., 2013). In east Antarctica moss growth rates over the last 50 yrs have been linked to changes in wind speed and temperature and their influence on water availability (Clarke et al., 2012). A contributing factor is that air temperatures have increased past the critical temperature at which successful sexual reproduction (seed set) can now take place, changing the dominant mode of reproduction, and increasing the potential distance for dispersal (Convey, 2011) (low confidence). Similar changes in the local distribution and development of typical cryptogamic vegetation of this region have been reported (Convey, 2011), including the rapid colonisation of ice free ground made available through glacial retreat and reduction in extent of previously permanent snow cover (Olech and Chwedorzewska, 2011). As these vegetation changes create new habitat, there are concurrent changes in the local distribution and abundance of the invertebrate fauna that then colonise them (low confidence). 28.2.4. Health and Well-Being of Arctic Residents The warming Arctic and major changes in the cryosphere are significantly impacting the health and well-being of Arctic residents and projected to increase especially, for many indigenous peoples. While impacts are expected to vary among the diverse settlements that range from small, remote predominantly indigenous to large cities and industrial settlements, this section focuses more on health impacts of climate change on indigenous, isolated, and rural populations because they are especially vulnerable to climate change due to a strong dependence on the Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 environment for food, culture and way of life; their political and economic marginalization; existing social, health, poverty disparities; as well as their frequent close proximity to exposed locations along ocean, lake or river shorelines (Ford and Furgal, 2009; Galloway-McLean, 2010; Larsen et al., 2010; Cochran et al., 2013). 28.2.4.1. Direct Impacts of a Changing Climate on the Health of Arctic Residents Direct impacts of climate changes on the health of Arctic residents include extreme weather events, rapidly changing weather conditions, and increasingly unsafe hunting conditions (physical/mental injuries, death, disease), temperature-related stress (limits of human survival in thermal environment, cold injuries, cold-related diseases), and UV-B radiation (immunosuppression, skin cancer, non-Hodgkin s lymphoma, cataracts) (Revich, 2008; AMAP, 2009 ; IPCC, 2012). Intense precipitation events and rapid snowmelt are expected to impact the magnitude and frequency of slumping and active layer detachment resulting in rock falls, debris flow, and avalanches (Kokelj et al., 2009; Ford et al., 2010). Other impacts from weather, extreme events, and natural disasters are the possibility of increasingly unpredictable, long duration and/or rapid onset of extreme weather events, storms, inundation by large storm surges, which, in turn, may create risks to safe travel or subsistence activities, loss of access to critical supplies and services to rural or isolated communities (e.g. food, fuel, telecommunications), and risk of being trapped outside one s own community (Laidre et al., 2008; Parkinson, 2009; Brubaker et al., 2011b). Changing river and sea ice conditions affect the safety of travel for indigenous populations especially, and inhibit access to critical hunting, herding and fishing areas (Andrachuk and Pearce, 2010; Derksen et al., 2012; Huntington and Watson, 2012). Cold exposure has been shown to increase the frequency of certain injuries (e.g. hypothermia, frostbite) or accidents, and diseases (respiratory, circulatory, cardiovascular, musculoskeletal skin) (Revich and Shaposhmikov, 2010). Studies in Northern Russia have indicated an association between low temperatures and social stress and cases of cardiomyopathy (Revich and Shaposhnikov, 2010). It is expected that winter warming in the Arctic will reduce winter mortality rates, primarily through a reduction in respiratory and cardiovascular deaths (Shaposhnikov et al., 2010). Researchers project that a reduction in cold-related injuries may occur, assuming that the standard for protection against the cold is not reduced (including individual behavior-related factors) (Nayha, 2005). Conversely, studies are showing respiratory and cardio stress associated with extreme warm summer days and that rising temperatures are accompanied by increased air pollution and mortality, especially in Russian cities with large pollution sources (Revich, 2008; Revich and Shaposhnikov, 2012). 28.2.4.2. Indirect Impacts of Climate Change on the Health of Arctic Residents Indirect effects of climate change on the health of Arctic residents include a complex set of impacts such as changes in animal and plant populations (species responses, infectious diseases), changes in the physical environment (ice and snow, permafrost), diet (food yields, availability of country food), the built environment (sanitation infrastructure, water supply system, waste systems, building structures), drinking water access, contaminants (local, long-range transported), and coastal issues (harmful algal blooms, erosion) (Maynard and Conway, 2007; Brubaker et al., 2011a; Parkinson and Evengard, 2009; Chapter 11). In addition to the climate change impacts and processes, are the complicated impacts from contaminants such as POPs (persistent organic pollutants), radioactivity, and heavy metals (e.g., mercury) which create additional and/or synergistic impacts on the overall health and well-being of all Arctic communities (Armitage et al., 2011; UNEP/AMAP, 2011; Teran et al., 2012). Ambient temperature variability and temperature gradients directly affect the volatilization, remobilization, and transport pathways of mercury and POPs in the atmosphere, ocean currents, sea ice and rivers. Transport pathways, intercompartmental distribution, bioaccumulation and transformation of environmental contaminants such as persistent organic pollutants (POPs), mercury (Hg) and radionuclides in the Arctic may consequently be affected by climate change (AMAP 2011;UNEP/AMAP 2011; Ma et al., 2011; Teng et al., 2012) (high confidence). Ma et al. (2011) and Hung et al. (2010) demonstrated that POPs are already being remobilized into the air from sinks in the Arctic region as a result of decreasing sea ice and increasing temperatures. Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Contaminants and human health in the Arctic are tightly linked to the climate and Arctic ecosystems by factors such as contaminant cycling and climate (increased transport to and from the Arctic), and the related increased risks of transmission to residents through subsistence life ways (Maynard, 2006; AMAP, 2010; Armitage et al., 2011; UNEP/AMAP 2011; Teran et al., 2012). The consumption of traditional foods by indigenous peoples places these populations at the top of the Arctic food chain and through biomagnification, therefore, they may receive some of the highest exposures in the world to certain contaminants (Armitage et al., 2011; UNEP/AMAP, 2011). Contaminants such as POPs are known for their adverse neurological and medical effects on humans, particularly, the developing fetus, children, women of reproductive age and the elderly, thus it is important to include contaminants as a significant part of any climate impact assessment (UNEP/AMAP, 2011). Radioactivity in the Arctic is also a concern because there are many potential and existing radionuclide sources in some parts of the Arctic and contamination can remain for long periods of time in soils and some vegetation, creating potentially high exposures for people (AMAP, 2010). Climate changes can mobilize radionuclides throughout the Arctic environment, and also potentially impact infrastructure associated with nuclear activities by changes in permafrost, precipitation, erosion, and extreme weather events (AMAP, 2010). Warming temperatures are enabling increased overwintering survival and distribution of new insects that sting and bite as well as many bird, animal, and insect species that can serve as disease vectors and, in turn, causing an increase in human exposure to new and emerging infectious diseases (Parkinson et al., 2008; Epstein and Ferber, 2011). Examples of new and emerging diseases are tick-borne encephalitis (brain infection) in Russia and Canada (Ogden et al., 2010; Tokarevich et al., 2011) and Sweden (Lindgren and Gustafson, 2001), Giardia spp. and Cryptosporidium spp. infection of ringed seals (Phoca hispida) and bowhead whales (Balaena mysticetus) in the Arctic Ocean (Hughes-Hanks et al., 2005). It is also expected that temperature increases will increase the incidence of zoonotic diseases as relocations of animal populations occur (Revich et al., 2012; Hueffler et al., 2013). Harmful algal blooms (HABs), whose biotoxins can be a serious health hazard to humans or animals (paralysis, death), are increasing globally and expected to increase in the Arctic, and HABS are influenced directly by climate change related factors such as temperature, winds, currents, nutrients and runoff (Portier et al., 2010; Epstein and Ferber, 2011; Walsh et al., 2011; Chapters 6; 11). Increasing ocean temperatures have caused an outbreak of a cholera-like disease, Vibrio parahaemolyticusin, in Alaskan oysters (McLaughlin et al., 2005). In addition, warmer temperatures raise the possibility of anthrax exposure in Siberia from permafrost thawing of historic cattle burial grounds (Revich and Podolnaya, 2011). The impacts of climate change on food security and basic nutrition are critical to human health because subsistence foods from the local environment provide Arctic residents, especially, indigenous peoples, with unique cultural and economic benefits necessary to well-being and contribute a significant proportion of daily requirements of nutrition, vitamins and essential elements to the diet (Ford and Berrang-Ford, 2009; Ford, 2009). However, climate change is already an important threat due to the decrease in predictability of weather patterns, low water levels and streams, timing of snow, ice extent and stability, impacting the opportunities for successful hunting, gathering, fishing and access to food sources and increasing the probability of accidents (Ford and Furgal, 2009; Ford et al., 2010). Populations of marine and land mammals, fish and water fowl are also being reduced or displaced, thus, reducing the traditional food supply (Gearheard et al., 2006; West and Hovelsrud, 2010; Lynn et al., 2013). Traditional food preservation methods such as drying of fish and meat, fermentation, and ice cellar storage are being compromised by warming temperatures, thus further reducing food available to the community (Brubaker et al., 2011b). For example, food contamination caused by thawing of permafrost ice cellars is occurring and increasingly wet conditions make it harder to dry food for storage (Hovelsrud et al., 2011). Indigenous people increasingly have to abandon their semi-nomadic lifestyles, limiting their overall flexibility to access traditional foods from more distant locations (www.arctichealthyukon.ca). These reductions in the availability of traditional foods plus general globalization pressures are forcing indigenous communities to increasingly depend upon expensive, non-traditional and often less healthy western foods, increasing the rates of modern diseases associated with processed food and its packaging, such as cardiovascular diseases, diabetes, dental cavities, and obesity (Armitage et al., 2011; Berrang-Ford et al., 2011; Brubaker et al., 2011b). Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climate change is beginning to threaten community and public health infrastructure, often in communities with no central water supply and treatment sources. This is especially serious in low-lying coastal Arctic communities (e.g., Shishmaref, Alaska, USA; Tuktoyaktuk, Northwest Territories, Canada) through increased river and coastal flooding and erosion, increased drought and thawing of permafrost, resulting in loss of reservoirs, damage to landfill sites, or sewage contamination (GAO, 2009; Bronen, 2011). Salt-water intrusion and bacterial contamination may also be threatening community water supplies (Parkinson et al., 2008; Virginia and Yalowitz, 2012). Quantities of water available for drinking, basic hygiene and cooking are becoming limited due to damaged infrastructure, drought, and changes in hydrology (Virginia and Yalowitz, 2012). Disease incidence caused by contact with human waste may increase when flooding and damaged infrastructure spreads sewage in villages with no municipal water supply. This can result in higher rates of hospitalization for pneumonia, influenza, skin infections, and respiratory viral infections (Parkinson and Evengard, 2009; Virginia and Yalowitz, 2012). Compounding these impacts in rural areas as well as cities are respiratory and other illnesses caused by airborne pollutants (e.g., contaminants, microbes, dust, mold, pollen, smoke) (Revich, 2008; Rylander and Schilling, 2011; Revich and Shaposhnikov, 2012). It is now well-documented that the many climate-related impacts on Arctic communities are causing significant psychological and mental distress and anxiety among residents (Portier et al., 2010; Coyle and Susteren, 2012; AR5 Chapter 11; Levintova, 2010). For example, changes in the physical environment (e.g., through thawing permafrost and erosion) which may lead to forced or voluntary relocation of residents out of their villages or loss of traditional subsistence species are causing mental health impacts among indigenous and other vulnerable, isolated populations (Curtis et al., 2005; Albrecht et al., 2007; Coyle and Susteren, 2012; Maldonado et al, 2013). Special concern has been expressed by many communities about the unusually high and increasing numbers of suicides in the Arctic especially among indigenous youth, and efforts are under way to try to develop a thorough assessment as well as establish effective intervention efforts (Albrecht et al., 2007; Portier et al., 2010; USARC, 2010). 28.2.5. Indigenous Peoples and Traditional Knowledge Indigenous populations in the Arctic the original Native inhabitants of the region are considered especially vulnerable to climate change, due to their close relationship with the environment and its natural resources for physical, social, and cultural well-being (Nuttall et al., 2005; Parkinson, 2009; Cochran et al., 2013). Arctic indigenous peoples are estimated to number between 400,000 and 1.3 million (Bogoyavlensky and Siggner, 2004; Galloway-McLean, 2010). According to the 2010 census data, there are 68,3 thousand indigenous people living in the Russian Arctic. These Arctic residents depend heavily on the region s terrestrial, marine and freshwater renewable resources, including fish, mammals, birds, and plants; however, the ability of indigenous peoples to maintain traditional livelihoods such as hunting, harvesting, and herding is increasingly being threatened by the unprecedented rate of climate change (Nakashima et al., 2012; Cochran et al., 2013). In habitats across the Arctic, climate changes are affecting these livelihoods through decreased sea ice thickness and extent, less predictable weather, severe storms, sea level rise, changing seasonal melt/freeze-up of rivers and lakes, changes in snow type and timing, increasing shrub growth, permafrost thaw, and storm-related erosion which, in turn, are causing such severe loss of land in some regions that a number of Alaskan coastal villages are having to relocate entire communities (Oskal, 2008; Mahoney et al., 2009; Forbes and Stammler, 2009; Bartsch et al., 2010; Weatherhead et al., 2010; Brubaker et al., 2011b; Bronen, 2011; Bongo et al., 2012; Eira et al., 2012; McNeeley, 2012; Huntington and Watson, 2012; Maldonado et al., 2013). In addressing these climate impacts, indigenous communities must at the same time consider multiple other stressors such as resource development (oil and gas, mining), pollution, changes in land use policies, changing forms of governance, and the prevalence in many indigenous communities of poverty, marginalization, and resulting health disparities (Abryutina, 2009; Reinert et al., 2009; Magga et al., 2011; Nakashima et al., 2012 Vuojala-Magga et al., 2011). Traditional knowledge is the historical knowledge of indigenous peoples accumulated over many generations and it is increasingly emerging as an important knowledge base for more comprehensively addressing the impacts of environmental and other changes as well as development of appropriate adaptation strategies for indigenous communities (IPCC AR4; IPCC AR5, chp 15; Oskal, 2008; Reinert et al., 2008; Wildcat, 2009; Nakashima et al., 2012; Magga et al., 2011; Vuojala-Magga et al., 2011; Vogesser et al, 2013) For example, Saami reindeer herders have specialized knowledge of dynamic snow conditions, which mediate access to forage on autumn, winter and Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 spring reindeer rangelands (Roturier and Roue, 2009; Eira et al., 2012; Vikhamar-Schuler et al., 2013) and traditional governance systems for relating to natural environments (Sara, 2013). Increasingly, traditional knowledge is being combined with western scientific knowledge to develop more sustainable adaptation strategies for all communities in the changing climate. For example, at Clyde River, Nunavut, Canada, Inuit experts and scientists both note that wind speed has increased in recent years and that wind direction changes more often over shorter periods (within a day) than it did during the past few decades (Gearheard et al., 2010; Overland et al., 2012). In Norway, Sámi reindeer herders and scientists are both observing direct and indirect impacts to reindeer husbandry such as changes in snow and ice cover, forage availability and timing of river freeze-thaw patterns from increasing temperatures. (Eira et al.,2012). On the Yamal Peninsula in western Siberia, detailed Nenets observations and recollections of iced-over autumn and winter pastures due to rain-on-snow events have proven suitable for calibrating the satellite-based microwave sensor SeaWinds (Bartsch et al., 2010) and NASA s AMSR-E sensor (Bongo et al., 2012). 28.2.6. Economic Sectors 28.2.6.1. Arctic 28.2.6.1.1. Agriculture and forestry Climate change presents benefits and costs for forestry and agriculture (Aaheim et al., 2009; Hovelsrud et al., 2011). In Iceland for example tree limits are found at higher altitudes than before, and productivity of many plants has increased (Björnsson et al., 2011). Grain production in Iceland, has increased in the last two decades, and work on soil conservation and forestry has benefited from warming (Sigurdsson et al., 2007; Björnsson et al., 2011), but also the number of new insect pests on trees and shrubs has increased in the past 20 years. A strong relationship between rate of new insect pest colonisation and outbrake intensity in forests exists with changes in annual temperature during the past century (Halldórsson et al., 2013). Climate change impacts on species change and fire frequency have potential impact on commercial forest harvesting activity. Vulnerability of forestry to changes that affect road conditions and thus accessibility during thawing periods has been found in Sweden (Keskitalo, 2008). A case study on Greenland found challenges for plant diseases in potatoes and grass fields, with pathogens and pests present in agricultural cropping systems, e.g. black scurf (Rhizoctonia) and common scab (Streptomyces scabies) (Neergaard et al., 2009). 28.2.6.1.2. Open and freshwater fisheries Current commercial fisheries are sharply divided between regions of high-yield and value commercial fisheries in the southern Bering Sea, Baffin Bay, the east and west Greenland Seas, the Iceland Shelf Sea, the deep Norwegian/Greenland Sea, and the Barents Seas and subsistence fisheries in the coastal regions of the Arctic Ocean. The relative absence of commercial fishing activity in the Arctic Ocean results from a combination of fisheries policy, the abundance of the resource, the lack of infrastructure for capturing and processing fish, and the difficulties in accessing fishing grounds especially during winter. In most regions, fisheries management strategies have been developed to build sustainable fisheries and rebuild overfished stocks (Froese and Proelß, 2010; Livingston et al., 2011). Recently observed changes in the spatial distribution and abundance of mackerel (Scomber scombrus) has challenged existing international agreements for shared resources in the North Atlantic (Arnason, 2012; Astthorsson et al., 2012). Although loss of sea ice in summer is allowing greater access to fisheries resources in the Arctic Ocean, some nations have prohibited commercial fishing within their EEZ until there is sufficient understanding of stock status to ensure that proposed fisheries would be managed sustainably (Wilson and Ormseth, 2009; Stram and Evans, 2009). Several Arctic coastal sea-run fishes are targeted for subsistence and commercial use in the Arctic. Commercial transactions from fishing are typically for local markets, however, the socioeconomic and cultural importance of these fishes to Indigenous Peoples far outweighs their monetary value. Reist et al. (2006) and Fechhelm et al. (2007) Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 found that climate related factors that influenced the water level and freshening of rivers were related to run size of arctic cisco (Coregonus autumnalis). Similarly, a recent study based on Chinook salmon (Oncorhynchus tshawytscha) run timing for the period (1961-2009) showed that success in the fishery was dependent on the timing of the marine exit, which was tightly coupled to environmental conditions that were linked to climate (Mundy and Evenson, 2011). 28.2.6.1.3. Marine transportation Observations and climate models indicate that in the period between 1979-1988 and 1998-2007 the number of days with ice free conditions (less than 15% ice concentration) increased by 22 days along the Northern Sea Route (NSR) in the Russian Arctic, and by 19 days in the Northwest Passage (NWP) in the Canadian Arctic, while the average duration of the navigation season in the period 1980-1999 was 45 and 35 days, respectively (Mokhow and Khon, 2008). Increased shipping associated with the opening of the NSR will lead to increased resource extraction on land and in the sea, and with two-way commodity flows between the Atlantic and Pacific. The future status of marine, terrestrial and freshwater biota may be negatively affected due to substantial coastal infrastructure to facilitate offshore developments (Meschtyb et al., 2010). Also, the frequency of marine transportation along the NSR is at its highest during the most productive and vulnerable season for fish and marine mammals, which is the late spring/summer, when these resources can be found throughout the NSR area (Ostreng, 2006). 28.2.6.1.4. Infrastructure Much of the physical infrastructure in the Arctic rely on and are adapted to local sea-ice conditions, permafrost, and snow (Huntington et al., 2007; Sundby and Nakken, 2008; West and Hovelsrud, 2010; Forbes, 2011; Sherman et al., 2009). Damage from ice action and flooding to installations such as bridges, pipelines, drilling platforms, and hydropower poses major economic costs and risks, which are more closely linked to the design of the structure than with thawing permafrost. Current engineering practices are designed to help minimize the impacts (Prowse et al., 2009). Much of the infrastructure has been built with weather conditions in mind, but remains vulnerable and inadequate to respond to environmental emergencies, natural disasters, and non-environmental accidents (NRTEE, 2009). Northern safety, security, and environmental integrity are much dependent upon transportation infrastructure. Ice as a provisioning system provides a transportation corridor and a platform for a range of activities and access to food sources in the Arctic (Eicken et al., 2009). In Northern Canada climate warming presents an additional challenge for northern development and infrastructure design. While the impacts of climate change become increasingly significant over the longer time scales, in the short term of greater significance will be the impacts associated with ground disturbance and construction (Smith and Risebrough, 2010). Climate change impacts have increased the demand for improved communication infrastructure and related services and community infrastructure for the safety and confidence in drinking water (NRTEE, 2009). The access, treatment and distribution of drinking water is generally dependent upon a stable platform of permafrost for pond or lake retention. Several communities have reported the need for more frequent water-quality testing both municipal systems and untreated water sources to ensure its suitability for drinking (Furgal, 2008). 28.2.6.1.5. Resource exploration The Arctic has large reserves of minerals (Lindholt, 2006; Peters et al., 2011; Harsem et al., 2011) and potentially large reserves of undiscovered sources of raw minerals, and oil and gas. Predicted new access to offshore energy resources is hypothesized to be a significant share of the global supply of oil and gas (Gautier et al., 2009; Berkman et al., 2010).The socio-economic impacts of oil and gas exploration activity may be positive or negative (Duhaime et al., 2004; Huntington et al., 2007; Forbes, 2008; Kumpula et al., 2011; Forbes et al., 2009; Harsem et al., 2011). Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Climatic warming is accelerating access to northern lands for development (Forbes et al., 2009). Yamal in Western Siberia has approximately 90 % of Russia s gas reserves, but at the same time represents the largest area of reindeer herding in the world (Jernsletten and Klokov, 2002; Stammler, 2005; Forbes and Kumpula, 2009). Development activities to obtain these resources would shrink the grazing lands, and have been characterized as one of the major human activities in the Arctic contributing to loss of available room for adaptation for reindeer husbandry (Oskal, 2008; Forbes et al., 2009; Nuttall et al., 2005). Sharp increases in future oil and gas and other resource development in the Russian North and other Arctic regions is anticipated - along with associated infrastructure, pollution, and other by products of development which will reduce the availability of pasturelands for reindeer and indigenous communities (Derome and Lukina, 2011; Degteva and Nellermann, 2013). 28.2.6.1.6. Informal, subsistence-based economy Hunting, gathering, herding, and fishing for subsistence, as well as commercial fishing, all play an important role in the mixed cash-subsistence economies (Nuttall et al., 2005; Poppel and Kruse, 2009; Larsen and Huskey, 2010; Crate et al., 2010). In the early 1990s initially in western Canada, and later elsewhere - indigenous communities started reporting climate change impacts (Berkes and Armitage, 2010). According to some herders, whalers and walrus hunters non-predictable conditions resulting from more frequent occurrence of unusual weather events are the main effect of recent warming (Forbes and Stammler, 2009; Ignatowski and Rosales, 2013; Forbes et al., 2009). The Inuit and Saami have expressed strong concern about the effects of climate warming on their livelihoods (Forbes and Stammler, 2009; Magga et al., 2011). For the Inuit, the issues revolve around sea ice conditions, such as later freeze-up in autumn, earlier melt-out and faster sea ice retreat in spring, and thinner, less predictable ice in general (Krupnik and Jolly, 2002; Cochran et al., 2013). Diminished sea ice translates into more difficult access for hunting marine mammals, and greater risk for the long-term viability of subsistence species such as polar bear populations (Laidre et al., 2008). Most Inuit communities depend to some extent on marine mammals for nutritional and cultural reasons, and many benefit economically from polar bear and narwhal hunting. A reduction in these resources represents a potentially significant economic loss (Hovelsrud et al., 2008). Among Fennoscandian Saami, the economic viability of reindeer herding is threatened by competition with other land users coupled with strict agricultural norms (Forbes, 2006; Magga et al., 2011). Reindeer herders are concerned that more extreme weather may exacerbate this situation (Oskal, 2008). Climate change is affecting reindeer herding communities through greater variability in snow melt/freeze, ice, weather, winds, temperatures and precipitation, which, in turn are affecting snow quality and quantity the most critical environmental variables for reindeer sustainability (Magga et al., 2011; Eira et al., 2012). Increasing temperature variations in wintertime, with temperatures rising above freezing with rain, followed by refreezing ( rain-on-snow conditions), are becoming more frequent, forming ice layers in the snow which then block the animals access to their forage and subsequent starvation (Maynard et al., 2011; Eira et al., 2012; Bongo et al., 2012). 28.2.6.2. Antarctica and the Southern Ocean Economic activities in the Antarctic have been limited to fishing and tourism (IPCC WG2, 2007). Ship-based tourism is a significant industry in Antarctica but does not involve permanent shore-based infrastructure. Over recent decades, the number of tourists landing in Antarctica has risen from 7322 in 1996/1997 to 32,637 in 2007/2008 (IAATO, 2012). Visits generally coincide with the times when wildlife are breeding and are often restricted because of the presence of fast ice, sea ice or icebergs. They are expected to continue to increase, with an increasing chance of terrestrial alien species being introduced from tourism and other vectors as ice-free areas increase from climate change (Chown et al., 2012). Scientific activity by a number of nations is also taking place and has the potential to impact upon local ecologies. Mineral resource activity is prohibited south of 60°S under the Protocol on Environmental Protection to the Antarctic Treaty. Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Fisheries in Antarctica, primarily through fisheries for Antarctic krill, could amount to approximately 6% of existing global marine capture fisheries (Nicol et al., 2011). The pattern of the krill fishery has been affected by changes in the sea ice extent around the Antarctic Peninsula where the fishery has been taking advantage of the ice-free conditions and taking more of its catch during winter in that region (high confidence) (Kawaguchi et al., 2009). Ecosystem-based management of krill fisheries by the Commission for the Conservation of Antarctic Marine Living Resources (CCAMLR) is yet to include procedures to account for climate change impacts, although the need to do so has been identified (Trathan and Agnew, 2010; Constable, 2011). 28.3. Key Projected Impacts and Vulnerabilities 28.3.1. Hydrology and Freshwater Ecosystems 28.3.1.1. Arctic Accompanying projected increases in high-latitude river flow (WGI Ch.12.4.5.4. and WGII Ch. 3.4.5.) are earlier spring runoff (Dankers and Middelkoop, 2008; Hay and McCabe, 2010; Pohl et al., 2007), greater spring snowmelt (Adam et al., 2009) and increases in spring sediment fluxes (Lewis and Lamoureux, 2010). Enhanced permafrost thaw (WGI Ch. 12.4.6.2) will continue to affect the dynamics of thermokarst lakes and related ecological effects (Section 28.2.1.1). Thawing permafrost and changes in the hydrological regime of the Arctic rivers, particularly those traversing regions affected by industrial developments, will increase the contaminant flow (Nikanorov et al., 2007). Loss of glacier ice masses will alter runoff hydrographs, sediment loads, water chemistry, thermal regimes, and related channel stability, habitat and biodiversity (Milner et al., 2009; Moore et al. 2009). Although snow, freshwater ice and permafrost affect the morphology of arctic alluvial channels, their future combined effects remain unclear (McNamara and Kane, 2009). For small permafrost streams, however, longer projected periods of flowing water will modify nutrient and organic matter processing (Greenwald et al., 2008; Zarnetske et al., 2008) but long-term negative impacts of increased sediment load on biological productivity could outweigh any positive effects from increased nutrient loading (Bowden et al., 2008). Changes to river-ice flooding are also projected to occur due to changes in i) hydraulic gradients for near-coastal locations because of sea-level rise, ii) streamwise air-temperature gradients, and iii) the timing and magnitude of spring snowmelt (Prowse et al. 2011). Synergistic/antagonistic effects among these factors, however, require detailed site-specific analyses for accurate projections of future conditions (Beltaos and Prowse, 2009). Reduced (increased) ice-jam flooding will have positive (negative) benefits for river-side northern communities/infrastructure but could also alter delta-riparian (Lesack and Marsh, 2010) and coastal-marine (Emmerton et al., 2008) ecosystems. The quality of river water entering the marine environment will also be affected by the reduction or loss of stamukhi lakes that process river inputs (Dumas et al., 2006; Galand et al., 2008). Future changes to lake-ice regimes will include: delayed freeze-up, advanced break-up, thinner ice and changes in cover composition (especially white ice in areas of enhanced winter precipitation), increased water temperature, and earlier and longer-lasting summer stratification (Dibike et al. 2011), all of which will affect a range of aquatic processes, including secondary productivity (Prowse and Brown, 2010b; Borgstrm and Museth, 2005; Prowse et al., 2007). Patterns of species richness and diversity are also projected to change with alterations to ice duration - increased open-water periods favouring the development of new trophic levels, colonization of new aquatic species assemblages (Vincent et al., 2009), greater atmosphere-water gas exchange and a decrease in winter kill of resident fish with cascading effects on lower trophic levels (Balayla et al., 2010). The loss of ice, however, can also decrease key habitat availability and quality (Vincent et al., 2008). Geochemical responses of Arctic lakes will also be altered. As observed for thermokarst lakes, the loss of ice cover and associated warming can greatly increase methane production (Laurion et al., 2010; Metje and Frenzel, 2007). Because temperature sensitivity has a stronger control over methane production than oxidation (Duc et al., 2010), elevated water temperatures will enhance methanogenesis, causing increased methane release from sediments. The Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 net balance of these two processes operating under a broad range of future changing environmental factors, however, remains to be quantified (Laurion et al., 2010; Walter et al., 2007a, b; 2008). As well as methane, increased water temperatures are projected to lead to reduced organic carbon (OC) burial. Projections, based on a range of six climate warming scenarios (Solomon et al., 2007), indicates that there will be a 4-27% decrease (0.9-6.4 TgC yr-1) in OC burial across lakes of the northern boreal zone by the end of the 21st Century as compared to rates for the approximately last half-century (Gudasz et al., 2010). Although these estimates assume that future organic carbon delivery will be similar to present-day conditions, even with enhanced supply from thawing permafrost, higher water temperatures will increase organic carbon mineralization and thereby lower burial efficiency. The amount of burial also depends on lake depth and mixing regimes. For non-thermally stratified shallow lakes, there will be a greater opportunity for water-sediment mixing and hence, greater carbon recycling back into the water column. By contrast, for lakes that become increasingly thermally stratified, carbon sinking below the thermocline will tend not to return to the surface until an increasing later fall turnover, thereby decreasing the probability of sediment-stored carbon being returned to the water column (Flanagan et al., 2006). Changes in ice cover, thermal regimes and stratification patterns will also affect the fate of contaminants in northern lakes. Higher water temperatures can enhance the methylation of mercury and modify food-web and energy pathways, such as through enhanced algal scavenging (a major foodweb entry pathway for mercury) resulting in increased mercury bio-availability to higher trophic levels (Carrie et al., 2010; Outridge et al., 2007). 28.3.1.2. Antarctica This assessment reinforces conclusions of AR4. Increased temperatures will impact aquatic ecosystems in Antarctica (high confidence) but the exact nature of these impacts will vary regionally. The most vulnerable freshwater systems are in the northern Antarctic Peninsula and maritime Antarctic islands, where a small increase in temperature can have widespread ecosystem impacts because the average temperature is within a few degrees of the melting point (Quesada and Velázquez 2012) (high confidence). Potential impacts are expected to range from immediate catastrophic impacts such as loss of bounding ice masses causing drainage of freshwater and epishelf lakes (Smith et al., 2006; Hodgson, 2011), to more gradual impacts on changes in the amount and duration of catchment ice and snow cover, accelerated glacier melting, declining volumes of precipitation falling as snow, permafrost, active layer and hydrological changes, such as water retention times (e.g. Vieira et al., 2010, Quesada and Velázquez 2012, Bockheim et al., 2013) (medium confidence). Changes in the thickness and duration of seasonal ice cover, longer melt seasons and larger volumes of water flowing into the lakes are expected in the future (Lyons et al., 2006) (medium confidence) but the ecological effects will vary between lakes, depending on their depth to surface area ratio, with insufficient evidence to fully assess future changes in these systems. Longer ice free seasons may cause physical conditions to be more favorable for primary production (Hodgson and Smol 2008) but very high irradiances experienced during summer in some systems can substantially inhibit algal blooms under ice free conditions (Tanabe et al., 2007), which would favor the growth of benthic cyanobacteria species (Hodgson et al., 2005). In other lakes, increases in meltwater supply may increase suspended solids and reduce light penetration and may offset the increases in the underwater light regime predicted as a result of extended ice free periods (Quesada et al., 2006). Under a warming climate an increase in microbial biomass is likely because of the increased water supply from glacial melt and warmer temperatures, and could result in further development of soils and elevated nutrient and dissolved organic carbon delivery to lakes (Velázquez et al., 2013). This organic supply will promote growth and reproduction in the benthos and plankton and imbalances in population dynamics (Quesada and Velázquez, 2013). Nutrient enrichment of some freshwater habitats in the vicinity of fur seal colonies will increase because of expanding fur seal populations (Quayle et al. 2013) (high confidence). Away from glacial forelands, increasing aridity will occur in the long-term in some areas of the continent (Hodgson et al., 2006b) and on subantarctic islands (Smith Jr et al., 2012) (medium confidence). Closed basin lakes can dry up completely causing local extinctions or retreat into cryptic or resistant life-cycle stages, as experienced in Arctic Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 lakes (Smol and Douglas, 2007b). Other effects include dessication of moss banks due to increased evaporation and sublimation rates (Wasley et al., 2006) (medium confidence). Studies have also shown that warming of once cold freshwater habitats in Antarctica will allow the sub- and maritime Antarctic species to re-invade and establish self- maintaining populations on the Antarctic continent, particularly where human vectors are involved (Barnes et al., 2006, Hodgson et al., 2006b) (medium confidence). For other organisms with lower dispersal capabilities there is increasing evidence of endemism, particularly in microbial groups (Vyverman et al., 2010), with a possibility that surface Antarctic lakes contain endemic species that are relicts of Gondwana (cf. Convey and Stevens 2007) and that would become extinct should they be lost from these lakes as a result of climate change.. 28.3.2. Oceanography and Marine Ecosystems 28.3.2.1. Ocean Acidification in the Arctic and Antarctic Ocean acidificaiton on polar marine food webs can have considerable implications (medium certainty). For example, if some regions in the Arctic become understaturated with respect to aragonite (the primary structural component of the shells of some marine calcifiers such as mollusks and urchins) the growth and survial of these organisms will be impacted (WG I, Chapter 6 Figure 6.28; Chierici and Fransson, 2009; Fabry et al., 2009; Yamamoto-Kawai et al., 2009). In laboratory experiments, Arctic pteropods (Limacina helicina, a small planktonic mollusk) held under conditions consistent with projected ocean warming and acidification in the Arctic Ocean in early spring were able to extend their shells in corosive waters but dissolution marks were observed (Comeau et al., 2010; Comeau et al., 2012). Additional studies are needed to scale up regional impacts to assess the population level impact of ocean acidification on Limacina helicina and other vulnerable species (Orr et al., 2009). At the current time there is insufficient data to fully assess the ecosystem consequences of acidification on pteropods because it is unclear whether other species, with a similar nutritive value, will replace pteropods. In the Southern Ocean, foraminifera have thinner shells than in the Holocene and there is evidence for shell thickness to be related to atmospheric CO2, supporting the hypothesis that ocean acidification will affect this abundant protozoan in this region (Moy et al., 2009). Similarly, shells are thinner from sediment traps in aragonite under-saturated water (below the aragonite saturation horizon - ASH) compared to those captured above the ASH in Subantarctic waters, but there is no time series of data related to change in the ASH (Roberts et al., 2011). Shell dissolution has been observed in surface waters in the Atlantic sector as a result of both upwelling and atmosperhic changes in CO2 (Bednarsek et al., 2012) (medium confidence). Other impacts of acidification on Southern Ocean organisms are currently uncertain, but short term negative impacts need to be considered together with an organism s capacity to adapt in the longer term (Watson et al., 2012). Only a few studies have been conducted on commercially exploited polar species on ocean acidification. Antarctic krill embryonic development (Kawaguchi et al., 2011) and post-larval krill metabolic physiology (Saba et al., 2012) may be impeded by elevated CO2 concentrations, which may negatively impact the reproductive success of krill more generally under emission scenarios used in CMIP5 (Kawaguchi et al., 2013) (medium confidence). Long et al. (2013) examined the effects of acidification on red king crab (Paralithodes camtschaticus) and found animals exposed to reduced pH exhibited increased hatch duration, decreased egg yolk, increased larval size, and decreased larval survival. In contrast, Hurst et al., (2012) conducted laboratory experiments at levels of elevated CO2 predicted to be present in the Gulf of Alaska and Bering Sea in the next century and found that juvenile walleye pollock exhibited a general resiliency of growth energetics to the direct effects of CO2 changes. Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 28.3.2.2. Arctic 28.3.2.2.1. Marine plankton, fish, and other invertebrates Phenological response Projected changes in the timing, spatial distribution and intensity of spring blooms may result in mis-matches with the timing of the emergence of Arctic grazers (Sreide et al., 2010). Based on past experience, some species will adapt to local conditions by shifting key life cycle events (hatch-date, maturity schedule and reproductive timing) or diet to accommodate differences in the regional timing and availability of prey and environmental conditions (Ormseth and Norcross, 2007; Sundby and Nakken, 2008; Vikeb et al., 2010, Darnis et al.,2012). For example, loss of sea ice cover in spring is expected to change fish behavior in ice bound areas (Mundy and Evenson, 2011). It is uncertain whether endemic animals will be able to alter key phenologies fast enough to keep pace with the projected rates of change in the Arctic Ocean. Projected spatial shifts Simulation studies revealed that a 2 week longer growing season and a 2 degree C increase in temperature would not be sufficient to allow expatriate species (Calanus finmarchicus or C. marshallae) to invade the Arctic Ocean (Ji et al., 2012). Ellingsen et al., (2008) projected future zooplankton distribution and abundance in the Barents Sea for the period 1995-2059 using a regional climate model which was forced with climate model output based on the IPCC-SRES B2 scenario. They projected that by 2059, Atlantic origin zooplankton will increase and Arctic origin zooplankton will decrease in the Barents Sea. The literature is mixed with respect to the potential for future movement of fish and shellfish into the Arctic Ocean. Modeling studies project that marine fish stocks potentially will shift their distributions into the Arctic Ocean resulting in an increase in biodiversity in the region (Cheung et al., 2009; Cheung et al., 2011; Box CC-MB). However, other studies show the persistence of cold sea water temperatures on the shelf regions of the Arctic Ocean and Northern Bering Sea will restrict or retard the movement of several sub-arctic fish and shellfish species into the Arctic Ocean (Sigler et al., 2011; Stabeno et al., 2012b; Hunt et al., 2013). In waters off the coasts of Europe there is a potential for increased fish production because of the combined effects of intrusion of Atlantic water over the relatively broader shelf regions and advective corridors for larval drift and range expansion of spawners. Huse and Ellingsen (2008) forced a spatially explicit coupled bio-physical model for the Barents Sea with future climate scenarios to project the implications of climate change on the spawning distribution of capelin (Mallotus villosus). Projections show that the spawning distribution of capelin will shift to the east and new spawning grounds will be colonized. A key factor governing this expansion will be the availability of pelagic prey. In the Bering Sea, there is evidence that planktivorous species like walleye pollock (Theragra chalcogramma) in the eastern Bering Sea will shift their distribution in response to shifts in ocean temperature (Kotwicki and Lauth, 2013). In summary, the spatial distribution of fish and shellfish in the Barents and Bering Seas will shift in response to climate change (high confidence) Projected impacts on production In the deep basins of the Arctic Ocean the number of ice free days in summer are expected to result longer productive seasons (Slagstad et al., 2011, high confidence). Ellingsen et al. (2008) projected that annual primary production would increase by 2059 in the Barents Sea. Tremblay et al. (2012) hypothesized that longer ice free periods in summer in the Arctic Ocean could provide for more opportunities for episodic nutrient pulses that would enhance secondary production through the growing season. However, in the Arctic Ocean, these changes in primary production may be offset later in the year by increased zooplankton grazing (Olli et al., 2007) or nutrient depletion due to stronger stratification and shifts in the mixed layer depth (Tremblay et al., 2012; Wassmann, 2011). Therefore, there is medium confidence that annual phytoplankton production will increase the central Arctic Ocean. Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In the few cases where future abundance of fish has been projected using climate change scenarios, species exhibited different trends related to their vulnerability. Forward extrapolation of observed responses suggests that increased summer sea surface temperatures in the Bering Sea and Barents Sea will cause a decrease in the abundance of energy rich copepods and euphausiids (Coyle et al., 2011; Slagstad et al., 2011). This change in prey quality is expected to lower survival of walleye pollock in the eastern Bering Sea by 2050 (Mueter et al., 2011). Climate enhanced stock projection models showed time trends in cross-shelf transport of juvenile northern rock sole (Lepidopsetta polyxystra) to nursery areas will not be substantially altered by climate change (Wilderbuer et al., 2012). 28.3.2.2.2. Marine mammals, polar bear, and seabirds The effects of the projected reduction is sea ice extent in this century (Wang and Overland, 2011) on Arctic marine mammals and sea birds will vary spatially and temporally (Laidre et al., 2008). Many ice-associated marine mammals and sea birds will be affected by ice loss with altered species distributions, migration patterns, behaviour, interspecific interactions, demography, population changes, and vulnerability to extinction but there is limited evidence of changes for most species (high confidence). The polar bear population of the Southern Beaufort Sea is projected to decline by 99% by 2100 with a probability estimated at 0.80-0.94 under A1B (Hunter et al., 2010). The Northern Beaufort Sea population is stable although decline is predicted with warming (Stirling et al., 2011). Projected extirpation of approximately two-thirds of the world s polar bears was predicted for mid-century under A1B (Amstrup et al., 2008). Aspects of this study were criticized (Armstrong et al., 2008) but refuted (Amstrup et al., 2009). The two-thirds decline is consistent with other studies and has robust evidence with medium agreement. Projected extinction of polar bears is unlikely. There is very high confidence of subpopulation extirpation. Several factors other than climate influence sea bird population dynamics (Regular et al., 2010), and projections of changes with a continued Arctic warming are therefore highly uncertain. Pattern of change will be non-uniform and highly complex (ACIA, 2005). At present, the resolution of AOGCMs are not detailed enough to project spatial changes in mesoscale oceanographic features like frontal zones and eddies of importance to Arctic sea birds. It is likely that the high Arctic seabird species partly or completely dependent on the sympagic ecosystem or the cold Arctic waters close to the ice-edge will be negatively impacted if the projected changes in these physical parameters occur (medium confidence). A general increase in SSTs, retreat of the ice cover, and earlier break up of fast ice may improve the environmental conditions and food abundance for sea bird species that have their range in the southern part of the Arctic or south of the Arctic (medium confidence). A poleward expansion of the range of these species is expected during a continued warming (medium confidence). Several factors other than climate influence sea bird population dynamics (Regular et al., 2010), and projections of changes with a continued Arctic warming are therefore highly uncertain. Pattern of change will be non-uniform and highly complex (ACIA 2005). At present, the resolution of AOGCMs are not detailed enough to project spatial changes in mesoscale oceanographic features like frontal zones and eddies of importance to Arctic sea birds. 28.3.2.3. Antarctica and the Southern Ocean Continued rising temperatures in the Southern Ocean will result in increased metabolic costs in many ectothermic pelagic species, southward movement of temperate species and contraction of the range of polar species (medium confidence). Southward movement of ocean fronts and associated biota that are prey of subantarctic island-based predators, will result in energetic inefficiencies for some of those predators (Péron et al., 2012; Weimerskirch et al., 2012) (low confidence). For Antarctic krill, insufficient evidence is available to predict what will happen to the circumpolar productivity of krill because of regional variability of the effects of climate change on the different factors (positive and negative) Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 that affect krill, directly and indirectly. For example, increased metabolic and growth rates from warming may be countered by a reduced food supply and the effects of ocean acidification (28.2.2.2, 28.3.2.1). Also, areas that are already warm may result in slower growth with further warming, such as could happen in the northern Scotia Arc (Wiedenmann et al., 2008; Hill et al., 2013). Models of recruitment and population dynamics indicate that the biomass of krill will decline if surface warming continues, but preliminary projections incorporating a range of factors are uncertain (low confidence) (Murphy et al., 2007, 2012b). Physiological and behavioural responses might also ameliorate impacts. For example, krill are now known to exploit the full depth of the ocean, which could provide escapes from further warming (Schmidt et al., 2011) as well as refuge from air-breathing predators. The strong dependence of species in more southern regions (e.g. southern west Antarctic Peninsula, WAP, and Ross Sea region) on sea-ice means that changes in sea-ice distribution will cause spatial shifts in the structure of ice- obligate food webs (Murphy et al., 2012b) (low confidence). Projections show that the loss of summer sea ice from the west Antarctic Peninsula is expected to result in ice-dependent seals declining and being replaced by other seal species that are not dependent on sea ice (Siniff et al., 2008, Costa et al., 2010) (low confidence). There is insufficient evidence to determine whether there will be a mismatch in phenologies of different species as a result of changes in the winter sea ice season (timing and winter extent), such as might occur if the timing of sea ice melt was not at a time of optimal growing conditions for phytoplankton (Trathan and Agnew, 2010). Reductions in krill abundance in the marine food webs around the South Atlantic islands may result in a shift in their structure towards a more fish-centred ecosystem as observed in the Indian Sector (Trathan, et al., 2007; Shreeve et al., 2009; Waluda et al., 2010; Trathan et al., 2012; Murphy et al., 2012a; Murphy et al., 2012b) (low confidence). Also, salps have been postulated to be competitors with krill for phytoplankton around the Antarctic Peninsula when oceanic conditions displace shelf and near-shelf waters during times of low sea ice (Ducklow et al., 2012). In the absence of krill, longer food chains have lower trophic efficiency (Murphy et al., 2012; Muprhy et al., 2013) and the long-term implications of this for higher trophic levels are unknown. Coastal environments will be impacted by the dynamics of fast ice, ice shelves and glacier tongues. These factors will positively affect local primary production and food web dynamics (Peck et al., 2009) but negatively affect benthic communities (Barnes and Souster, 2011) (low confidence). Projections of the response of emperor penguins and Southern Ocean seabirds based on AR4 model outputs for sea ice and temperature in east Antarctica indicate that general declines in these populations are to be expected if sea ice habitats decline in the future (Barbraud et al., 2011; Jenouvrier et al., 2012) (low confidence). However, these responses are also expected to be regionally specific because of the regional differences in expectations of change in the ice habitats (high confidence). Additional studies at other sites are needed to improve confidence levels of predictions. 28.3.3. Terrestrial Environment and Related Ecosystems 28.3.3.1. Arctic The boreal forest is generally projected by models to move northward under a warming climate, that will displace between 11% and 50% of the tundra within 100 years (Callaghan et al., 2005; Wolf et al., 2008; Tchebakova et al., 2009; Wramneby et al., 2010 in a pattern similar to that which occurred during the early Holocene climatic warming (high confidence). Pearson et al. (2013) projected that at least half of vegetated Arctic areas will shift to a different physiognomic class, and woody cover will increase by as much as 52%, in line with what has been occurring in northwest Eurasia (Macias-Fauria et al. 2012). Dynamic vegetation models applied to Europe and the Barents Region project a general increase in net annual primary production by climate warming and CO2 fertilization (Wolf et al., 2008; Wramneby et al., 2010; Anisimov et al., 2011). Boreal needle-leaved evergreen coniferous forest replaces tundra and expands into the mountain areas of Fennoscandia, but this advance may be delayed or prevented in regions already occupied by clonal deciduous shrubs whose in situ growth has increased significantly in recent decades (Macias-Fauria et al. 2012). Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In contrast to these expected results, shrubs, currently expanding in area in many Arctic locations, were modelled to decrease in extent over the next 100 years after an initial increase (Wolf et al., 2008). Also counter-intuitively, tundra areas increased in the projections. This was a result of changes at the highest latitudes that opened land for colonisation at a rate exceeding displacement of tundra by shrubs in the south. Several studies have calculated the magnitude of the effects of vegetation change in the Arctic on negative feedbacks of CO2 sequestration and increased evapo-transpiration and the positive feedback of decreased albedo (Swann et al.,2010; Wramneby et al., 2010; Wolf et al., 2010; Pearson et al., 2013). It is likely that vegetation changes will result in an overall positive feedback on the climate. Recent changes and results of climate change simulation experiments in the field have shown that there are considerable uncertainties in the projected rates of change e.g. (Van Bogaert et al. 2010). Furthermore, the models do not yet include vertebrate and invertebrate herbivory, extreme events such as tundra fire and extreme winter warming damage or changes in land use that either reduce the rate of vegetation change or open up niches for rapid change. Projections suggest increases in the ranges of the autumn and winter moths that have outbreaks in populations resulting in the defoliation of birch forest (Jepsen et al., 2008 and 2011), and a general increase in the background (non-outbreak) invertebrate herbivores (Wolf et al., 2008). Animal terrestrial biodiversity is generally projected to increase in the Arctic during warming by immigration of new species from the south, vegetation changes, and indirectly by introduction of invasive species caused by increased human activities and increased survival of such species (high confidence) (Post et al., 2009; Gilg et al., 2012; CAFF 2013). Many native Arctic species will most likely be increasingly threatened during this century. 28.3.3.2. Antarctica Projected effects of climate change on Antarctic terrestrial species are limited to knowledge of their ecophysiological tolerances to changes in air temperature, wind speed, precipitation (rain and snowfall), permafrost thaw and exposure of new habitat through glacial/ice retreat. The climate is expected to become more tolerable to a number of species, leading to increases in biomass and extent of existing ecological communities. The frequency with which new potential colonising plant and animal species arrive in Antarctica (particularly the Antarctic Peninsula region) from lower latitudes, and the subsequent probability of their successful establishment will increase with regional climate warming and associated environmental changes (Chown et al., 2012) (high confidence). Human-assisted transfers of biota may be more important by two orders of magnitude than natural introductions (Frenot et al., 2005) as the transfer is faster and avoids extreme environments such as altitude or oceans (Barnes et al., 2006). The potential for anthropogenic introduction of non-indigenous species to Antarctic terrestrial areas, which could have devastating consequences to the local biodiversity, will increase (Convey et al., 2009; Convey, 2011, Hughes and Convey, 2010; Braun et al., 2012) (high confidence). At present, established non- indigenous species in the sub- and maritime Antarctic are very restricted in their distributions (Frenot et al., 2005). Climate change could result in a greater rate of spread of invasive species through colonisation of areas exposed by glacial retreat, as has occurred at South Georgia (Cook et al., 2010) and in the maritime Antarctic (Olech and Chwedorzewska 2011). Biosecurity measures may be needed to help control dispersal of established non-indigenous species to new locations, particularly given the expected increase in human activities in terrestrial areas (Hughes and Convey, 2010; Convey et al., 2011). An important gap in understanding is the degree to which climate change may facilitate some established but localised alien species to become invasive and widespread (Frenot et al., 2005; Convey 2010; Hughes and Convey 2010; Cowan et al., 2011), which has been shown for the sub-Antarctic (Chown et al., 2012). Overall, the likely impacts of existing and new non-indigenous species on the native terrestrial ecosystems of Antarctica and the sub-Antarctic islands, along with the continued increased presence of Antarctic fur seals are likely to have far greater importance over the timescale under consideration than are those attributable to climate change itself (Turner et al., 2009; Convey and Lebouvier, 2009; Convey, 2010). Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 28.3.4. Economic Sectors Projections of economic costs of climate change impacts for different economic sectors in the Arctic are limited, but current assessments suggest that there will be both benefits and costs (Forbes, 2011; AMAP, 2011). Non-Arctic actors are likely to receive most of the benefits from increased shipping and commercial development of renewable and non-renewable resources, while indigenous peoples and local Arctic communities will have a harder time maintaining their way of life (Hovelsrud et al., 2011). Contributing to the complexity of measuring the future economic effects of climate change is the uncertainty in future predictions and the rapid speed of change, which are linked with the uncertainty of the technological and ecological effects of such change (NorAcia, 2010). Communities within the same eco-zone may experience different effects from identical climate-related events because of marked local variations in site, situation, culture and economy (Clark et al., 2008). Economic cost estimates have been made for the case of the Alaskan economy, for example, which suggest that a heavy reliance on climate-sensitive businesses such as tourism, forestry, and fisheries, renders the economy vulnerable to climate change, and that Alaska Native peoples, reliant on the biodiversity of the Alaskan ecosystem, are being affected disproportionately (Epstein and Ferber, 2011). Some Alaskan villages such as Shishmaref, Kivalina, and Newtok have already lost critical infrastructure and services and are becoming unlivable due to storm damage and coastal erosion but the high costs and limitations of government mechanisms are significant barriers to the actual relocation of these communities (Bronen, 2011; Cochran et al., 2013; Maldonado et al., 2013). 28.3.4.1. Fisheries Climate change will impact the spatial distribution and catch of some open ocean fisheries in the Barents and Bering Seas (high certainty). The future of commercial fisheries in Arctic Ocean is uncertain. There is strong evidence and considerable data showing links between climate driven shifts in ocean conditions and shifts in the spatial distribution and abundance of commercial species in the Bering and Barents Seas (Section 28.3.2.2.1). In limited cases, coupled bio-physical models or climate enhanced stock projection models have been used to predict future commercial yield or shifts in fishing locations. However, these predictions are uncertain (Huse and Ellingsen, 2008; Ianelli et al., 2011; Wilderbuer et al. 2012). Cheung et al. (2011) used projections from an earth system model to estimate shifts in bio-climatic windows that included climate change effects on biogeochemistry (oxygen and acidity) and primary production to project future catch potential of 120 demersal fish and invertebrates. Results from their model suggested that the catch potential will increase in the Barents and Greenland Seas and regions > 70o north latitude (Cheung et al., 2011). In contrast, vulnerability analysis suggests that only a few species are expected to be abundant enough to support viable fisheries in the Arctic Ocean (Hollowed et al., 2013). Potential fisheries for snow crab on shelf areas of the Arctic Ocean may be limited by the associated impacts of ocean acidification. If fisheries develop in the Arctic Ocean, adoption of sustainable strategies for management will be a high priority (Molenaar, 2009). The moratorium on fishing in the U.S. portion of the Chukchi and Beaufort Seas would prevent fishing until sufficient data become available to manage the stock sustainably (Wilson and Ormseth, 2009). Predicting of how harvesters will respond to changing economic, institutional and environmental conditions under climate change is difficult. Current techniques, track fishers choices based on revenues and costs associated with targeting a species in a given time and area with a particular gear given projected changes in the abundance and spatial distribution of target species (Haynie and Pfeiffer, 2012). However, estimates of future revenues and costs will depend, in part, on future demand for fish, global fish markets and trends in aquaculture practices (Merino et al., 2012; Rice and Garcia, 2011). Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 28.3.4.2. Forestry and Farming Climate change is likely to have positive impacts for agriculture, including extended growing season (Grnlund, 2009; Falloon and Betts, 2009; Tholstrup and Rasmussen, 2009), although variations across regions are expected (Hovelsrud et al., 2011), and the importance of impacts to the Arctic economy will likely remain minor (Eskeland and Flottorp, 2006). Potential positive effects of climatic warming for forestry include decreased risk of snow damage. Kilpelainen et al. (2010) estimate a 50% decrease in snow damage in Finland towards the end of the century. A warmer climate is likely to impact access conditions and plant diseases for forestry and farming. Grnlund (2009) found in the case of Northern Norway where about half of the arable land area is covered by forest and 40% by marshland that the potential harnessing of arable land for farming will be at the cost of forestry production, or dried-up marshlands, which may contribute to more greenhouse emissions. Larger field areas may contribute to land erosion through rainfall and predicted unstable winters, and may increase conditions for plant diseases and fungal infections (Grnlund, 2009). If the winter season continues to shorten due to climate change (Xu et al., 2013), accessibility to logging sites will be negatively affected. Accessibility is higher when frozen ground makes transportation possible in sensitive locations or areas that lack road. If weather changes occur when logging has taken place, sanding of roads may be necessary which carries significant economic costs. Impact on carrying capacity of ground or road accessibility will thus affect forestry economically. Challenges may include limited storage space for wood (Keskitalo, 2008). 28.3.4.3. Infrastructure, Transportation, and Terrestrial Resources Rising temperatures and changing precipitation patterns have the potential to affect all infrastructure types and related services, as much of the infrastructure in the North is dependent upon the cryosphere to, for example, provide stable surfaces for buildings and pipelines, contain waste, stabilize shorelines and provide access to remote communities in the winter (Furgal and Prowse, 2008; Huntington et al., 2007; Sherman et al., 2009; Sundby and Nakken, 2008; West and Hovelsrud, 2010; Forbes, 2011). In the long term marine and freshwater transportation will need to shift its reliance from ice routes to open-water or land-based transportation systems. Of appropriate community adaptations to the predicted changes relocation is one option to deal with persistent flooding and bank erosion (Furgal, 2008; NRTEE, 2009). Changing sea-ice (multiyear) conditions are suspected i.e. to have a regulating impact on marine shipping and coastal infrastructure through possible hazards on them (Eicken et. al., 2009). By adapting transportation models to integrate monthly climate model (CCSM3) predictions of air temperature, combined with datasets on land cover, topography, hydrography, built infrastructure, and locations of human settlements, estimates have been made of changes to inland accessibility for northern landscapes northward of 40N by mid-21st Century (Stephenson et al., 2011). Milder air temperatures and/or increased snowfall reduce the possibilities for constructing inland winter-road networks, including ice roads, with the major seasonal reductions in road potential (based on a 2000 kg vehicle) being in the winter shoulder-season months of November and April. The average decline (compared to a baseline of 2000-2014) for eight circumpolar countries was projected to be -14%, varying from -11 to -82%. In absolute terms, Canada and Russia (both at -13%) account for the majority of declining winter-road potential with ~1x106 km2 being lost (See Table 28.1). The winter road season has decreased since the 1970s on the Alaska´s North Slope, from as much as 200 to 100 days in some areas (Hinzman, et al., 2005). Climate change is expected to lead to a nearly ice free Arctic Ocean in late summer and increased navigability of Arctic marine waters. New possibilities for shipping routes and extended use of existing routes may result from increased melting of sea ice (Paxian et al., 2010; Corbett et al., 2010; Khon et al., 2010; Peters et al., 2011; Stephenson et al., 2011). Projections made by Stephenson et al. (2011) suggest that all five Arctic littoral states will gain increased maritime access to their current exclusive economic zones, especially Greenland (+28% relative to baseline), Canada (+19%), Russia (+16%) and United States (+15%). In contrast, Iceland, Norway, Sweden, and Finland display little or no increase in maritime accessibility,(Table 28.1) (Stephenson et al., 2011). Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 GCMs generally underestimate the duration of the ice-free period in the Arctic Ocean and simulate slower changes than those observed in the past decades (Stroeve et al., 2007). Mokhow and Khon (2008) used a sub-set of climate models that better than other GCMs reproduce the observed sea ice dynamics to project the duration of the navigation season along the NSR and through the NWP under the moderate SRES-A1B emission scenario. According to their results, by the end of the 21st century the NSR may be open for navigation 4.5+/-1.3 months per year, while the NWP may be open 2-4 months per year (Figure 28-4). The models did not predict any significant changes of the ice conditions in the NWP until the early 2030s. [INSERT FIGURE 28-4 HERE Figure 28-4: Projected duration of the navigation period (days) over the Northwest Passage and Northern Sea Route. Source: Mokhow and Khon, 2008.] [INSERT TABLE 28-1 HERE Table 28-1: Annually averaged changes in inland and maritime transportation accessibility by mid-century (2045 2059) versus baseline (2000 2014). Source: Stephenson et al., 2011.] An increase in the length of the summer shipping season, with sea-ice duration expected to be 10 days shorter by 2020 and 20-30 days shorter by 2080, is likely to be the most obvious impact of changing climate on Arctic marine transportation (Prowse et al., 2009). Reduction in sea ice and increased marine traffic could offer opportunities for economic diversification in new service sectors supporting marine shipping. Loss of sea ice may open up waterways and opportunities for increased cruise traffic (e.g. Glomsrd and Aslaksen, 2009), and add to an already rapid increase in cruise tourism (Stewart et al., 2010; Stewart et al., 2007; Howell et al., 2007). Climate change has increased the prevalence of cruise tourism throughout Greenland, Norway, Alaska and Canada because of decreasing sea ice extent. Projected declines in sea-ice covers leading to development of integrated land and marine transportation networks in Northern Canada may stimulate further mine exploration and development (Prowse et al., 2009). These possibilities however also come with challenges including their predicted contribution to the largest change in contaminant movement into or within the Arctic, as well as their significant negative impacts on the traditional ways of life of northern residents (Furgal and Prowse, 2008). Added shipping and economic activity will increase the amount of black carbon and reinforce warming trends in the region (Lack and Corbett, 2012), leading to additional economic activity. Longer shipping season and improved access to ports may lead to increased petroleum activities, although possible increased wave activity and coastal erosion may increase costs related to infrastructure and technology. Peters et al. (2011) find by using a bottom-up shipping model and a detailed global energy market model to construct emission inventories of Arctic shipping and petroleum activities in 2030 and 2050 and based on estimated sea-ice extent that there will be rapid growth in transit shipping; oil and gas production will be moving into locations requiring more ship transport; and this will lead to rapid growth in emissions from oil and gas transport by ship. The Arctic contains vast resources of oil, which is hard to replace as transportation fuel, and vast resources of gas, a more climate benign fuel than coal. Petroleum resources are unevenly distributed among Arctic regions and states. Arctic resources will play a growing role in the world economy, but increased accessibility is expected to create challenges for extraction, transport, engineering, search-and-rescue needs and responses to accidents (Hovelsrud et al., 2011), and climatic change presents the oil and gas industry with challenges in terms of planning and predictions (Harsem et al., 2011). Increased emissions due to rapid growth in Arctic Ocean transportation of oil and gas are projected (Peters et al., 2011). Due to high costs and difficult access conditions the impact on future oil and gas production in the Arctic remains unclear (Peters et al., 2011; Lindholdt and Glomsrd, 2012). Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 28.4. Human Adaptation There is general agreement that both indigenous and non-indigenous people in the Arctic have a history of adapting to natural variability in the climate and natural resource base, as well as recent socio-economic, cultural and technological changes (Forbes and Stammler, 2009; Wenzel, 2009; West and Hovelsrud, 2010; Ford and Pearce, 2010; Bolton et al., 2011Cochran et al., 2013). Climate change exacerbates the existing stresses faced by Arctic communities (Rybraten and Hovelsrud, 2010; Crate and Nuttall, 2009) and is only one of many important factors influencing adaptation (Berrang-Ford et al., 2011). Climate adaptation needs to be seen in the context of these inter- connected and mutually reinforcing factors (Tyler et al., 2007; Hovelsrud and Smit, 2010). The challenges faced today by communities in the Arctic are complex and interlinked and are testing their traditional adaptive capacity. Climatic and other large-scale changes have potentially large effects on Arctic communities, in particular where simple economies leave a narrower range of adaptive choices (Anisimov and Vaughan, 2007; Ford and Furgal, 2009; Andrachuk and Pearce, 2010; Ford et al., 2010; Forbes, 2011; Berkes et al., 2003). There is considerable evidence that changing weather patterns, declining sea-ice and river as well as lake ice, thawing permafrost, and plant and animal species abundance and composition have consequences for communities in the Arctic (see 28.2.4, 28.2.5.2 and 28.3.4). Sea-ice is particularly important for coastal communities which rely upon it for transportation between communities and hunting areas (Krupnik et al., 2010). Changes in the duration and condition of sea ice and the consequent changes to country food availability significantly impact the wellbeing of communities (Furgal and Seguin, 2006; Ford and Berrang-Ford, 2009; Ford et al., 2010), outdoor tourism (Dawson et al., 2010) and hunting and fishing (Wiig et al., 2008; Brander, 2010). Adaptation to climate change is taking place at the local and regional levels where impacts are often felt most acutely and the resources most readily available (Oskal, 2008; Hovelsrud and Smit, 2010). Current experiences and projections of future conditions often lead to technological adaptation responses such as flood and water management and snow avalanche protection (West and Hovelsrud, 2010; Hovelsrud and Smit, 2010) rather than policy responses (Hedensted Lund et al., 2012; Rudberg et al., 2012). Climate variability and extreme events are found to be salient drivers of adaptation (Berrang-Ford et al., 2011; Dannevig et al., 2012; Amundsen et al., 2010). The lack of local scale climate projections, combined with uncertainties in future economic, social and technological developments often act as barriers to adaptation. These barriers, together with other societal determinants such as ethics, cultures, and attitudes towards risk may cause inaction (Adger et al., 2009; West and Hovelsrud, 2010). Resolving divergent values across and within different communities poses a challenge for governance regimes. A determining factor in building adaptive capacity is the flexibility of enabling institutions to develop robust options (Keskitalo et al., 2009; Hovelsrud and Smit 2010; Forbes et al., 2009; Ford and Goldhar, 2012; Whyte, 2013). In the North American and Scandinavian context, adaptive co-management responses have been developed through land claims settlements and/or multi-scale institutional cooperation to foster social learning (Berkes, 2009; Armitage et al., 2008). Indigenous Peoples While Arctic indigenous peoples with traditional lifestyles are facing unprecedented impacts to their ways of life from climate change and resource development (oil and gas, mining, forestry, hydropower, tourism, etc.), they are already implementing creative ways of adapting (Cruikshank, 2001; Forbes et al., 2006; Krupnik and Ray, 2007; Salick and Ross, 2009; Green and Raygorodetsky, 2010; Cullen-Unsworth et al., 2011; Alexander et al., 2011; Bongo et al., 2012). Examples of indigenous adaptation strategies have included changing resource bases, shifting land use and/or settlement areas, combining technologies with traditional knowledge, changing timing and location of hunting, gathering, herding, and fishing areas, and improving communications and education (Galloway McLean, 2010; Bongo et al., 2012). Protection of grazing land will be the most important adaptive strategy for reindeer herders under climate change (Forbes et al., 2009; Magga et al., 2011; Kumpula et al., 2012; Degteva and Nellemann, 2013). Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The adaptive capacity of Arctic indigenous peoples is largely due to an extensive traditional knowledge and cultural repertoire, and flexible social networks (see Chapter 12, section 12.3) (Williams and Hardison, 2013). The dynamic nature of traditional knowledge is valuable for adapting to current conditions (Kitti et al.,. 2006; Tyler et al., 2007; Eira et al., 2012). The sharing of knowledge ensures rapid responses to crises (Ford et. al., 2007). In addition, cultural values such as sharing, patience, persistence, calmness, respect for elders and the environment are important. Some studies suggest that traditional knowledge may not always be sufficient to meet the rapid changes in climate (see also Chapter 12) and it may be perceived to be less reliable because the changing conditions are beyond the current knowledge range (Ingram et al., 2002; Ford et al., 2006; Valdivia et al., 2010; Hovelsrud et al., 2010). Over the last half-century, the adaptive capacity in some indigenous communities has been challenged by the transition from semi-nomadic hunting groups to permanent settlements (Ford et al., 2010). Forced or voluntary migration as an adaptation response can have deep cultural impacts (Shearer, 2011,2012; Maldonado et al., 2013).The establishment of permanent communities, particularly those associated with new industrial development, can also lead to increasing employment opportunities and income diversification for indigenous peoples. The intergenerational transfers of knowledge and skills through school curricula, land camps, and involvement in community-based monitoring programmes may strengthen adaptive capacity (Bolton et al., 2011; Hovelsrud and Smit, 2010; Ford et. al., 2007; Forbes 2007). Renewable resource harvesting remains a significant component of Arctic livelihoods and with climate change hunting and fishing has become a riskier undertaking and many communities are already adapting (Gearheard et al., 2011; Laidler et al., 2011). Adaptation includes taking more supplies when hunting; constructing permanent shelters on land as refuges from storms; improved communications infrastructure; greater use of global positioning systems (GPS) for navigation; synthetic aperture radar (SAR) to provide estimates of sea-ice conditions (Laidler et al., 2011) and the use of larger or faster vehicles (Ford et al., 2010). Avoiding dangerous terrain can result in longer and time- consuming journeys which can be inconvenient to those with wage-earning employment (Ford et al., 2007). Reindeer herders have developed a wide range of adaptation strategies in response to changing pasture conditions. These include: moving herds to better pastures (Bartsch et al., 2010); providing supplemental feeding (Helle and Jaakkola, 2008; Forbes and Kumpula, 2009); retaining a few castrated reindeer males to break through heavy ice- crust (Oskal, 2008; Reinert et al., 2008); ensuring an optimal herd size (Forbes et al., 2009; Tyler et al., 2007); and creating multicultural initiatives combining traditional knowledge with scientific (Vuojala-Magga et al., 2011; Bongo et al., 2012). Coastal fishers have adapted to changing climate by targeting different species and diversifying income sources (Hovelsrud et al., 2010). In some Arctic countries indigenous peoples have successfully negotiated land claims rights and have become key players in addressing climate change (Abele et al., 2009). In some instances this has given rise to tensions over land/water use between traditional livelihoods and new opportunities (e.g. tourism and natural resource development) (Forbes et al., 2006; Hovelsrud and Smit, 2010). Some territorial governments in Northern Canada have promoted adaptation by providing hunter support programs (Ford et al., 2006, 2010). The health of many indigenous people is being affected by the interaction of changes in the climate with ongoing changes in human, economic and biophysical systems (Donaldson et al., 2010). The distribution of traditional foods between communities and the use of community freezers in the Canadian Arctic has improved food security, an important factor for health (Ford et al., 2010). While wage employment may enhance the possibilities for adaptive capacity, greater involvement in full time jobs can threaten social and cultural cohesion and mental well-being by disrupting the traditional cycle of land-based practices (Berner et al., 2005; Furgal, 2008). 28.5. Research and Data Gaps There remains a poor knowledge of coupling among, and thresholds within, bio-geophysical and socio-economic processes to fully assess the effects of a changing climate, and to separate them from those due to other environmental stressors: Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Existing integrative models are either lacking or insufficiently validated to project and to assess the cascading effects on, and feedbacks from the systems in the Polar Regions, in particular socio-economic systems. There is a need to enhance or establish a coordinated network of long-term representative sites for monitoring and assessment of climate change detection and attribution studies in Polar Regions. Regional differences and confounding variables will need to be considered in designing field and modelling studies. Standardised methods and approaches of biophysical and socio-economic analysis along with coordinated sampling in more regions will be necessary. There are more specific research gaps, including: Many mechanisms of how climate change and ocean acidification may be affecting polar ecosystems have been proposed but few studies of physiological tolerances of species, long term field studies of ecosystem effects and ecosystem modelling studies are available to be able to attribute with high confidence current and future change in these ecosystems to climate change. More comprehensive studies including long-term monitoring on the increasing impacts from climate changes on Arctic communities (urban and rural) and their health, well-being, traditional livelihoods and life ways are needed. There is a need to assess more fully vulnerabilities and to develop response capacities at the local and regional level. Frequently Asked Questions FAQ 28.1: What will be the net socio-economic impacts of change in the polar regions? [to remain at the end of the chapter] Climate change will have costs and benefits for Polar Regions. Climate change, exacerbated by other large-scale changes, can have potentially large effects on Arctic communities, where relatively simple economies leave a narrower range of adaptive choices. In the Arctic, positive impacts include new possibilities for economic diversification, marine shipping, agricultural production, forestry, and tourism. The Northern Sea Route is predicted to have up to 125 days per year suitable for navigation by 2050, while the heating energy demand in the populated Arctic areas is predicted to decline by 15%. In addition, there could be greater accessibility to offshore mineral and energy resources although challenges related to environmental impacts and traditional livelihoods are possible. Changing sea ice condition and permafrost thawing may cause damage to bridges, pipelines, drilling platforms, hydropower and other infrastructure. This poses major economic costs and human risks, although these impacts are closely linked to the design of the structure. Furthermore, warmer winter temperatures will shorten the accessibility of ice roads that are critical for communications between settlements and economic development and have implications for increased costs.. Statistically, a long-term mean increase of 2 to 3°C in autumn and spring air temperature produces an approximate 10 to 15 day delay in freeze-up and advance in break-up, respectively. Particular concerns are associated with projected increase in the frequency and severity of ice-jam floods on Siberian rivers. They may have potentially catastrophic consequences for the villages and cities located in the river plain, as exemplified by the 2001 Lena River flood, which demolished most of the buildings in the city of Lensk. Changing sea ice conditions will impact indigenous livelihoods, and changes in resources, including marine mammals, could represent a significant economic loss for many local communities. Food security and health and well-being are expected to be impacted negatively. In the Antarctic, tourism is expected to increase, and risks exist of accidental pollution from maritime accidents, along with an increasing likelihood of the introduction of alien species to terrestrial environments. Fishing for Antarctic krill near to the Antarctic continent is expected to become more common during winter months in areas where there is less winter sea ice. FAQ 28.2: Why are changes in sea ice so important to the polar regions? [to remain at the end of the chapter] Sea ice is a dominant feature of Polar Oceans. Shifts in the distribution and extent of sea ice during the growing season impacts the duration, magnitude and species composition of primary and secondary production in the Polar Regions. With less sea ice many marine ecosystems will experience more light, which can accelerate the growth of phytoplankton, and shift the balance between the primary production by ice algae and water-borne phytoplankton, Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 with implications for Arctic food webs. In contrast, sea ice is also an important habitat for juvenile Antarctic krill, providing food and protection from predators. Krill is a basic food source for many species in polar marine ecosystems. Changes in sea ice will have other impacts, beyond these bottom-up consequences for marine foodwebs. Mammals and birds utilize sea ice as haul-outs during foraging trips (seals, walrus, and polar bears in the Arctic and seals and penguins in the Antarctic). Some seals (e.g. Bearded seals in the Arctic and crabeater and leopard seals in the Antarctic) give birth and nurse pups in pack ice. Shifts in the spatial distribution and extent of sea ice will alter the spatial overlap of predators and their prey. According to model projections, within 50-70 years loss of hunting habitats may lead to elimination of polar bears from seasonally ice-covered areas, where two thirds of their world population currently live. The vulnerability of marine species to changes in sea ice will depend on the exposure to change, which will vary by location, as well as the sensitivity of the species to changing environmental conditions and the adaptive capacity of each species. More open waters and longer ice-free period in the northern seas enhance the effect of wave action and coastal erosion with implications for coastal communities and infrastructure. While the overall sea ice extent in the Southern Ocean has not changed markedly in recent decades, there have been increases in oceanic temperatures and large regional decreases in winter sea ice extent and duration in the western Antarctic Peninsula region of West Antarctica and the islands of the Scotia Arc. References Aaheim, A., H. Dannevig, T. Ericsson, B. van Oort, L. Innbjr, T. Rauken, H. Vennemo, H. Johansen, M. Tofteng, C. Aall, K. Groven and E. Heiberg, 2009. Konsekvenser av klimaendringer, tilpasning og sarbarhet i Norge Rapport til Klimatilpasningsutvalget. [in Norwegian]. CICERO Report. 4:238 pp. Abele, F., T. J. Courchene, F.L. Sidele, and F. St-Hilaire, 2009: Northern Exposure: Peoples, Powers and Prospects in Canada's North. Montreal: The Institute for Research on Public Policy. xii, 605 p. : ill., Montréal, Québec Abryutina, L.I., 2009: Indigenous peoples of the Russian North: Social and Climatic changes. In: Climate Change and Arctic Sustainable Development: Scientific, Social Cultural, and educational challenges. UNESCO, Paris, pp. 164-173. Adam, J.C., A.F. Hamlet, and D.P. Lettenmaier, 2009: Implications of global climate change for snowmelt hydrology in the twenty-first century. Hydrological Processes, 23(7), 962-972. Adger, W., S. Dessai, M. Goulden, M. Hulme, I. Lorenzoni, D. Nelson, L. Naess, J. Wolf, and A. Wreford, 2009: Are there social limits to adaptation to climate change? Climatic Change, 93(3), 335-354. Albrecht, G., G. Sartore, L. Connor, N. Higginbotham, S. Freeman, B. Kelly, H. Stain, A. Tonna, and G. Pollard, 2007: Solastalgia: the distress caused by environmental change. Australasian Psychiatry, 15, S95-S98. Alexander, C., N. Bynum, E. Johnson, U. King, T. Mustonen, P. Neofotis, N. Oettlé, C. Rosenzweig, C. Sakakibara, V. Shadrin, M. Vicarelli, Jon Waterhouse, and B. Weeks, 2011: Linking Indigenous and Scientific Knowledge of Climate Change. Bioscience, 61(6), pp. 477-484. AMAP, 2011. Arctic Pollution 2011. Arctic Monitoring and Assessment Programme (AMAP). Oslo, Norway. Vi + 38 pp. AMAP, 2011: Snow, Water, Ice and Permafrost in the Arctic (SWIPA): Climate Change and the Cryosphere. Arctic Monitoring and Assessment Programme (AMAP), Oslo, Norway. xii + 538 pp. AMAP. 2009. AMAP Assessment 2009: Human Health in the Arctic. Arctic Monitoring and Assessment Programme (AMAP). Oslo. Norway. Xii + 254 pp. AMAP. 2010. AMAP Assessment 2009: Radioactivity in the Arctic. Arctic Monitoring and Assessment Programme (AMAP). Oslo, Norway. Xii + 92. Amstrup, S.C., B.G. Marcot, and D.C. Douglas, 2008: A Bayesian network approach to forecasting the 21st century worldwide status of polar bears. In: Arctic Sea Ice Decline: Observations, Projections, Mechanisms, and Implications. [DeWeaver, E.T., C.M. Bitz, and L.B. Tremblay (eds.)]. AGU, Washington, D. C., pp. 213. Amstrup, S.C., E.T. DeWeaver, D.C. Douglas, B.G. Marcot, G.M. Durner, C.M. Bitz, and D.A. Bailey, 2010: Greenhouse gas mitigation can reduce sea-ice loss and increase polar bear persistence. Nature, 468(7326), 955- U351 Amstrup, S.C., H. Caswell, E. DeWeaver, I. Stirling, D.C. Douglas, B.G. Marcot, and C.M. Hunter, 2009: Rebuttal of "Polar Bear Population Forecasts: A Public-Policy Forecasting Audit". Interfaces, doi:10.1287/inte.1090.0444. Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Amstrup, S.C., I. Stirling, T.S. Smith, C. Perham, and G.W. Thiemann, 2006: Recent observations of intraspecific predation and cannibalism among polar bears in the southern Beaufort Sea. Polar Biology, 29(11), 997-1002. Amundsen, Helene, Frode Viktor Berglund and Hege Westskog, 2010. Overcoming barriers to climate change adaptation a question of multilevel governance?. Environment and Planning C, 28 (2): pp. 276-289. Andrachuk, M. and T. Pearce, 2010: Vulnerability and Adaptation in Two Communities in the Inuvialuit Settlement Region. In: Community Adaptation and Vulnerability in Arctic Regions. [Hovelsrud, G.K. and B. Smit(eds.)]. Springer Netherlands, pp. 63-81. Anisimov, O. A., Zhil'tsova, E. L. and Reneva, S. A. 2011. Estimation of critical levels of climate change influence on the natural terrestrial ecosystems on the territory of Russia. Meteorology and Hydrology 36, 723-730 DOI: 10.3103/S1068373911110033 Anisimov, O. and D. Vaughan, 2007: Polar regions. In: Climate Change 2007:Impacts, adaptation, and vulnerability. Contribution of Working group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, pp. 653-686. Arctic Council, 2013. Arctic Resilience Interim Report 2013. Stockholm Environment Institute and Stockholm Resilience Centre, Stockholm. Armitage D., 2008: Governance and the commons in a multi-level world. International Journal of the Commons, 2(1) Armitage, J.M., C.L. Quinn, and F. Wania. 2011. Global Climate change and contaminants an overview of opportunities and priorities for modeling the potential implications for long-term human exposure to organic compounds in the Arctic. J. Environ. Monit. 13, 1532. Pp. 1532-1546. Armstrong, J.S., K.C. Green, and W. Soon, 2008: Polar Bear Population Forecasts: A Public-Policy Forecasting Audit. Interfaces, 38(5), 382-395. Arnason, R., 2012: Global warming: New challenges for the common fisheries policy? Ocean & Coastal Management, 70, 4-9. Arrigo, K., G. van Dijken, and S. Bushinsky, 2008: Primary production in the Southern Ocean, 1997-2006. Journal of Geophysical Research-Oceans, 113, [C8]. Arrigo, K.R., and G. L. van Dijken, 2011: Secular trends in Arctic Ocean net primary production. Journal of Geophysical Research, 116, C09011. Arrigo, K.R., D.K. Perovich, R.S. Pickart, Z.W. Brown, G.L. van Dijken, K.E. Lowry, M.M. Mills, M.A. Palmer, W.M. Balch, F. Bahr, N.R. Bates, C. Benitez-Nelson, B. Bowler, E. Brownlee, J.K. Ehn, K.E. Frey, R. Garley, S.R. Laney, L. Lubelczyk, J. Mathis, A. Matsuoka, B.G. Mitchell, G.W.K. Moore, E. Ortega-Retuerta, S. Pal, C.M. Polashenski, R.A. Reynolds, B. Schieber, H.M. Sosik, M. Stephens, and J.H. Swift, 2012: Massive Phytoplankton Blooms Under Arctic Sea Ice. Science, 336(6087), 1408-1408. Astthorsson, O. S., H. Valdimarsson, A. Gudmundsdottir, G. J. O´skarsson. 2013: Climate-related variations in the occurrence and distribution of mackerel (Scomber scombrus) in Icelandic waters. ICES Journal of Marine Science, 69(7), 1289 1297. Atkinson, A., V. Siegel, E. Pakhomov, and P. Rothery, 2004: Long-term decline in krill stock and increase in salps within the Southern Ocean. Nature, 2996(Letters to Nature), 1-4. Bajzak, C.E., M.O. Hammill, G.B. Stenson, and S. Prinsenberg, 2011: Drifting away: Implications of changes in ice conditions for a pack-ice-breeding phocid, the harp seal (Pagophilus groenlandicus). Canadian Journal of Zoology, 89(11), 1050-1062. Bakun, A., 2010: Linking climate to population variability in marine ecosystems characterized by non-simple dynamics: Conceptual templates and schematic constructs. Journal of Marine Systems, 79(3 4), 361-373. Balayla, D., T. Lauridsen, M. SA¸ndergaard, and E. Jeppesen, 2010: Larger zooplankton in Danish lakes after cold winters: are winter fish kills of importance? Hydrobiologia, 646(1), 159-172. Barber, D.G., J.V. Lukovich, J. Keogak, S. Baryluk, L. Fortier, and G.H.R. Henry, 2008: The Changing Climate of the Arctic. Arctic, 61, 7-26. Barbraud, C., C. Marteau, V. Ridoux, K. Delord, and H. Weimerskirch, 2008: Demographic response of a population of white-chinned petrels Procellaria aequinoctialis to climate and longline fishery bycatch. Journal of Applied Ecology, 45(5), 1460-1467. Barbraud, C., P. Rivalan, P. Inchausti, M. Nevoux, V. Rolland, and H. Weimerskirch, 2011: Contrasted demographic responses facing future climate change in Southern Ocean seabirds. Journal of Animal Ecology, 80(1), 89-100. Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Barbraud, C., V. Rolland, S. Jenouvrier, M. Nevoux, K. Delord, and H. Weimerskirch, 2012: Effects of climate change and fisheries bycatch on Southern Ocean seabirds: a review. Marine Ecology Progress Series, 454, 285- 307. Barnes, D.K.A. and T. Souster, 2011: Reduced survival of Antarctic benthos linked to climate-induced iceberg scouring. Nature Climate Change, 1(7), 365-368. Barnes, D.K.A., D.A. Hodgson, P. Convey, C.S. Allen, and A. Clarke, 2006: Incursion and excursion of Antarctic biota: past, present and future. Global Ecology and Biogeography, 15(2), 121-142. Barnes, D.K.A., D.A. Hodgson, P. Convey, C.S. Allen, and A. Clarke, 2006: Incursion and excursion of Antarctic biota: past, present and future. Global Ecology and Biogeography, 15(2), 121-142 Bartsch, A., T. Kumpula, B.C. Forbes, and F. Stammler, 2010: Detection of snow surface thawing and refreezing in the Eurasian Arctic with QuikSCAT: implications for reindeer herding. Ecological Applications, 20(8), 2346- 2358. Beaugrand, G. and R.R. Kirby, 2010: Spatial changes in the sensitivity of Atlantic cod to climate-driven effects in the plankton. Climate Research, 41(1), 15-19. Bednarsek, N., G.A. Tarling, D.C.E. Bakker, S. Fielding, E.M. Jones, H.J. Venables, P. Ward, A. Kuzirian, B. Leze, R.A. Feely, and E.J. Murphy, 2012: Extensive dissolution of live pteropods in the Southern Ocean. Nature Geoscience, 5(12), 881-885. Beltaos, S. and T.D. Prowse, 2009. River ice hydrology in a shrinking cryosphere. Hydrological Processes, 23(1): 122-144. Berg, T.B., N.M. Schmidt, T.T. Hye, P.J. Aastrup, D.K. Hendrichsen, M.C. Forchhammer, and D.R. Klein, 2008: High-Arctic Plant-Herbivore Interactions under Climate Influence. Advances in Ecological Research, 40, 275- 298. Berg, T.B., N.M. Schmidt., T.T. Hye., P.J. Aastrup., D.K. Hendrichsen., M.C. Forchhammer, and D.R. Klein, 2008: High-Arctic Plant-Herbivore Interactions Under Climate Influence. Advances in Ecological Research, 40, 275-298. Berkes, F. and D. Armitage, 2010: Co-Management Institutions, Knowledge and Learning: Adapting to Change in the Arctic. Etudes Inuit Studies, 34(1), 109. Berkes, F., 2009: Evolution of co-management: Role of knowledge generation, bridging organizations and social learning. Journal of Environmental Management, 90(5), 1692-1702. Berkman, P.A. and Royal United Services Institute for Defence and Security Studies., 2010: Environmental security in the Arctic Ocean : promoting co-operation and preventing conflict. Published on behalf of The Royal United Services Institute for Defence and Security Studies by Routledge Journals, Abingdon, . Berrang-Ford, L., J.D. Ford, and J. Paterson, 2011: Are we adapting to climate change? Global Environmental Change, 21(1), 25-33. Berrang-Ford, L., J.D. Ford, and J. Paterson, 2011: Are we adapting to climate change? Global Environmental Change, 21(1), 25-33. Bhatt, U.S., D.A. Walker, M.K. Raynolds, J.C. Comiso, H.E. Epstein, G. Jia, R. Gens, J.E. Pinzon, C.J. Tucker, C.E. Tweedie, and P.J. Webber, 2010: Circumpolar Arctic tundra vegetation change is linked to sea ice decline. Earth Interactions, 14(8). Björk, R.G. and U. Molau, 2007: Ecology of alpine snowbeds and the impact of global change. Arctic, Antarctic, and Alpine Research, 39(1), 34-43. Björnsson H., T. Johannesson and A. Snorrason, 2011. Recent climate change, projected impacts and adaptation capacity in Iceland. Í: Linkov, I. & T. S. Bridges (Ed.) Global change and local adaptation. NATO Science for Peace and Security Series - C: Environmental Security. Springer, Dordrecht, s. 465-475 Bluhm, B.A. and R. Gradinger, 2008: Regional Variability in Food Availability for Arctic Marine Mammals. Ecological Applications, 18(2, Supplement: Arctic Marine Mammals and Climate Change), pp. S77-S96. Bockheim, J., G. Vieira, M. Ramos, J. Lopez-Martinez, E. Serrano, M. Guglielmin, K. Wilhelm, A. Nieuwendam, 2013: Climate Warming and Permafrost Dynamics in the Antarctic Peninsula Region . Global and Planetary Change, 100: 215-223 Bogoyavlensky, D. and A. Siggner, 2004: Arctic Demography. Arctic Human Development Report. Stefansson Arctic Institute, Akureyri, Iceland, pp. 27. Bokhorst, S., J.W. Bjerke, L.E. Street, T.V. Callaghan, and G.K. Phoenix, 2011: Impacts of multiple extreme winter warming events on sub-Arctic heathland: Phenology, reproduction, growth, and CO2 flux responses. Global Change Biology, 17(9), 2817-2830. Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bokhorst, S.F., J.W. Bjerke, H. Tmmervik, T.V. Callaghan, and G.K. Phoenix, 2009: Winter warming events damage sub-Arctic vegetation: Consistent evidence from an experimental manipulation and a natural event. Journal of Ecology, 97(6), 1408-1415. Bolton, K., M. Lougheed, J. Ford, S. Nickels, C. Grable, and J. Shirley, 2011: What we know, don't know, and need to know about climate change in Nunavut, Nunavik, and Nunatsiavut: A systematic literature review and gap analysis. Inuit Tapiriit Kanatami, Ottawa, Canada, pp. 128. Bongo, M.P., A. Degteva, I.M.G. Eira, I. Hanssen-Bauer, A. Ivanoff, O.H. Magga, S.D. Mathiesen, N.G. Maynard, A. Oskal, M. Pogodaev, M.N. Sara, D.V. Schuler, E.I. Turi, and R.W. Corell, 2012: Eurasian Reindeer Herding, Traditional Knowledge and Adaptation to Climate Change. In: UNESCO.Indigenous Peoples and Climate Change: Vulnerability, Adaptation and Traditional Knowledge. Report of workshop on Indigenous Peoples, Marginalized Populations and Climate Change: Vulnerability, Adaptation and Traditional Knowledge. UNESCO, Paris. Bongo, M.P., A. Degteva, I.M.G. Eira, I. Hanssen-Bauer, A. Ivanoff, O.H. Magga, S.D. Mathiesen, N.G. Maynard, A. Oskal, M. Pogodaev, M.N. Sara, D.V. Schuler, E.I. Turi, and R.W. Corell, 2012: Eurasian Reindeer Herding, Traditional Knowledge and Adaptation to Climate Change. In: UNESCO.Indigenous Peoples and Climate Change: Vulnerability, Adaptation and Traditional Knowledge. Report of workshop on Indigenous Peoples, Marginalized Populations and Climate Change: Vulnerability, Adaptation and Traditional Knowledge. UNESCO, Paris. Bongo, M.P., A. Degteva, I.M.G. Eira, I. Hanssen-Bauer, A. Ivanoff, O.H. Magga, S.D. Mathiesen, N.G. Maynard, A. Oskal, M. Pogodaev, M.N. Sara, D.V. Schuler, E.I. Turi, and R.W. Corell, 2012: Eurasian Reindeer Herding, Traditional Knowledge and Adaptation to Climate Change. In: UNESCO.Indigenous Peoples and Climate Change: Vulnerability, Adaptation and Traditional Knowledge. Report of workshop on Indigenous Peoples, Marginalized Populations and Climate Change: Vulnerability, Adaptation and Traditional Knowledge. UNESCO, Paris. Borgstrm, R. and J. Museth, 2005: Accumulated snow and summer temperature ? critical factors for recruitment to high mountain populations of brown trout (Salmo trutta L.). Ecology of Freshwater Fish, 14(4), 375-384. Bouchard, C. and L. Fortier, 2011: Circum-arctic comparison of the hatching season of polar cod Boreogadus saida: A test of the freshwater winter refuge hypothesis. Progress in Oceanography, 90(1-4), 105-116. Bowden, W.B., M.N. Gooseff, A. Balser, A. Green, B.J. Peterson, and J. Bradford, 2008: Sediment and nutrient delivery from thermokarst features in the foothills of the North Slope, Alaska: Potential impacts on headwater stream ecosystems. 113, G02026. Boyd, P.W., K.R. Arrigo, R. Strzepek, and G.L. van Dijken, 2012: Mapping phytoplankton iron utilization: Insights into Southern Ocean supply mechanisms. Journal of Geophysical Research: Oceans, 117(C6), - C06009. Brander, K.M., 2010: Cod Gadus morhua and climate change: processes, productivity and prediction. Journal of Fish Biology, 77, 1899. Braun, C., O. Mustafa, A. Nordt, S. Pfeiffer, and H. Peter, 2012: Environmental monitoring and management proposals for the Fildes region, King George Island, Antarctic. Polar Research, 31, http://dx.doi.org/10.3402/polar.v31i0.18206. Briffa, K.R., V.V. Shishov, T.M. Melvin, E.A. Vaganov, H. Grudd, R.M. Hantemirov, M. Eronen, and M.M. Naurzbaev, 2008: Trends in recent temperature and radial tree growth spanning 2000 years across northwest Eurasia. Philosophical Transactions of the Royal Society B: Biological Sciences, 363(1501), 2271-2284. Brommer, J.E., H. Pietiäinen, K. Ahola, P. Karell, T. Karstinenz, and H. Kolunen, 2010: The return of the vole cycle in southern Finland refutes the generality of the loss of cycles through 'climatic forcing'. Global Change Biology, 16(2), 577-586. Bronen, R. 2009. Forced Migration of Alaskan Indigenous Communities Due to Climate Change: Creating a Human Rights ResponseIn Linking Environmental Change, Migration & Social Vulnerability. A. Oliver-Smith and X. Shen, eds. Pp. 68-73. UNU Institute for Environment and Human Security. Hohenkammer, Germany. Bronen, R. 2011. Climate-induced community relocations: creating an adaptive governance framework based in human rights doctrine. NYU Review of Law and Social Change 35:356 406 Brown, Z. W., K. R. Arrigo, 2013: Sea ice impacts on spring bloom dynamics and net primary production in the Eastern Bering Sea. Journal of Geophysical Research: Oceans, 118, 43-62. Brubaker, M., J. Berner, R. Chavan, and J. Warren, 2011a: Climate change and health effects in Northwest Alaska. Global Health Action, 4, 10.3402/gha.v4i0.8445. Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Brubaker, M.Y., J.N. Bell, J.E. Berner, and J.A. Warren, 2011b: Climate change health assessment: a novel approach for Alaska Native communities. International Journal of Circumpolar Health, 70(3), 266-273. Brubaker, M.Y., J.N. Bell, J.E. Berner, and J.A. Warren, 2011b: Climate change health assessment: a novel approach for Alaska Native communities. International Journal of Circumpolar Health, 70(3), 266-273. Burek, K.A., F.M.D. Gulland, and T.M. O'Hara, 2008: effects of climate change on arctic marine mammal health. Ecological Applications, 18(2), S126-S134. Byrd, G.V., J.A. Schmutz, and H.M. Renner, 2008: Contrasting population trends of piscivorous seabirds in the Pribilof Islands: A 30-year perspective. Deep-Sea Research Part Ii-Topical Studies in Oceanography, 55(16- 17), 1846-1855. CAFF 2013. Arctic Biodiversity Assessment. Status and trends in Arctic biodiversity. Conservation of Arctic Flora and Fauna, Akureyri. 557pp Callaghan, T.V., C.E. Tweedie, J. Akerman, C. Andrews, J. Bergstedt, M.G. Butler, T.R. Christensen, D. Cooley, U. Dahlberg, R.K. Danby, F.J.A. Daniëls, J.G. De Molenaar, J. Dick, C.E. Mortensen, D. Ebert-May, U. Emanuelsson, H. Eriksson, H. Hedenas, G.H.R. Henry, D.S. Hik, J.E. Hobbie, E.J. Jantze, C. Jaspers, C. Johansson, M. Johansson, D.R. Johnson, J.F. Johnstone, C. Jonasson, C. Kennedy, A.J. Kenney, F. Keuper, S. Koh, C.J. Krebs, H. Lantuit, M.J. Lara, D. Lin, V.L. Lougheed, J. Madsen, N. Matveyeva, D.C. Mcewen, I.H. Myers-Smith, Y.K. Narozhniy, H. Olsson, V.A. Pohjola, L.W. Price, F. Rigét, S. Rundqvist, A. Sandström, M. Tamstorf, R.V. Bogaert, S. Villarreal, P.J. Webber, and V.A. Zemtsov, 2011b: Multi-decadal changes in tundra environments and ecosystems: Synthesis of the international Polar year-back to the future project (IPY-BTF). Ambio, 40(6), 705-716. Callaghan, T.V., F. Bergholm, T.R. Christensen, C. Jonasson, U. Kokfelt, and M. Johansson, 2010: A new climate era in the sub-Arctic: Accelerating climate changes and multiple impacts. Geophysical Research Letters, 37(14), 6. Callaghan, T.V., L.O. Björn, Y. Chernov, F.S. Chapin, T.R. Christensen, B. Huntley, R. Ims, S. Jonasson, D. Jolly, N. Matveyeva, N. Panikov, W.C. Oechel, G.R. Shaver, J. Elster, H. Henttonen, I.S. Jónsdóttir, K. Laine, S. Schaphoff, S. Sitch, E. Taulavuori, K. Taulavuori, and C. Zöckler, 2005: Tundra and Polar Desert Ecosystems. In: ACIA. Arctic Climate Impacts Assessment. [Symon, C., L. Arris, and B. Heal(eds.)]. Cambridge University Press, Cambridge, UK, pp. 243-352. Callaghan, T.V., M. Johansson, R.D. Brown, P.Y. Groisman, N. Labba, and V. Radionov, 2011c: Changing snow cover and its impacts. In: Snow, Water, Ice and Permafrost in the Arctic (SWIPA). AMAP, Oslo, Norway, pp. 4- 1 to 4-58. Callaghan, T.V., T.R. Christensen, and E.J. Jantze, 2011a: Plant and vegetation dynamics on Disko island, west Greenland: Snapshots separated by over 40 years. Ambio, 40(6), 624-637. Callaghan,T.V., C. Jonasson, T. Thierfelder, Z. Yang, H. Hedenas, M. Johansson, U. Molau, R. Van Bogaert, A. Michelsen, J. Olofsson, D. Gwynn-Jones, S. Bokhorst, G. Phoenix, J. W. Bjerke, H. Tommervik, T.R. Christensen, E. Hanna, E. K. Koller, V.L. Sloan, 2013: Ecosystem change and stability over multiple decades in the Swedish subarctic: complex processes and multiple drivers. Transactions of the Royal Society, 368(1624), http://dx.doi.org/10.1098/rstb.2012.0488 Canada NRTEE (National Round Table on the Environment and the Economy). 2009. True North: Adapting Infrastructure to Climate Change in Northern Canada. NRTEE, Ottawa, 146 p. (http://www.nrtee- trnee.com/eng/publications/true-north/true-north-eng.pdf). Carrie, J., F. Wang, H. Sanei, R.W. Macdonald, P.M. Outridge, and G.A. Stern, 2010: Increasing Contaminant Burdens in an Arctic Fish, Burbot (Lota lota), in a Warming Climate. Environmental Science & Technology, 44(1), 316-322. Castro de la Guardia, L., Derocher, A.E., Myers, P.G., Terwisscha van Scheltinga, A.D., and Lunn, N.J. 2013. Future sea ice conditions in western Hudson Bay and consequences for polar bears in the 21 century. Global Change Biology, 19, 2675-2687. Chapman, E.W., E.E. Hofmann, D.L. Patterson, C.A. Ribic, and W.R. Fraser, 2011: Marine and terrestrial factors affecting Adélie penguinPygoscelis adeliaechick growth and recruitment off the western Antarctic Peninsula. Marine Ecology Progress Series, 436, 273-289. Cherosov, M.M., A.P. Isaev, V.I. Mironova, L.P. Lytkina, L.D. Gavrilyeva, R.R. Sofronov, A.P. Arzhakova, N.V. Barashkova, I.A. Ivanov, I.F. Shurduk, A.P. Efimova, N.S. Karpov, P.A. Timofeyev, and L.V. Kuznetsova, 2010: Vegetation and human activity. In: The Far North: plant biodiversity and ecology of Yakutia. [Troeva, E.I., A.P. Isaev, M.M. Cherosov, and N.S. Karpov(eds.)]. Springer, Berlin, Germany, pp. 261-295. Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cherry, S.G., A.E. Derocher, I. Stirling, and E.S. Richardson, 2009: Fasting physiology of polar bears in relation to environmental change and breeding behavior in the Beaufort Sea. Polar Biology, 32(3), 383-391. Cheung, W. W. L., J. Dunne, J. L. Sarmienot, and D. Pauly, 2011: Integrating ecophysiology and plankton dynamics into projected maximum fisheries catch potential under climate change in the Northeast Atlantic. ICES Journal of Marine Fisheries, 68(6):1008-1018. Cheung, W.W.L., V.W.Y. Lam, J.L. Sarmiento, K. Kearney, R. Watson, and D. Pauly, 2009: Projecting global marine biodiversity impacts under climate change scenarios. Fish and Fisheries, 10(3), 235-251. Chierici, M. and A. Fransson, 2009: Calcium carbonate saturation in the surface water of the Arctic Ocean: undersaturation in freshwater influenced shelves. Biogeosciences, 6(11), 2421-2431. Chown, S.L., J.E. Lee, K.A. Hughes, J. Barmes, P.J. Barrett, D.M. Bergstrom, P. Convey, D.A. Cowan, K. Crosbie, G. Dyer, Y. Frenot, S.M. Grant, D. Herr, M.C.I. Kennicutt, M. Lamers, A. Murray, H.P. Possingham, K. Reid, M.J. Riddle, P.G. Ryan, L. Sanson, J.D. Shaw, M.D. Sparrow, C. Summerhayes, A. Terauds, and D.H. Wall, 2012: Challenges to the future conservation of the Antarctic. Science, 337, 158-159. Chown, S.L., J.E. Lee, K.A. Hughes, J. Barmes, P.J. Barrett, D.M. Bergstrom, P. Convey, D.A. Cowan, K. Crosbie, G. Dyer, Y. Frenot, S.M. Grant, D. Herr, M.C.I. Kennicutt, M. Lamers, A. Murray, H.P. Possingham, K. Reid, M.J. Riddle, P.G. Ryan, L. Sanson, J.D. Shaw, M.D. Sparrow, C. Summerhayes, A. Terauds, and D.H. Wall, 2012: Challenges to the future conservation of the Antarctic. Science, 337, 158-159. Christoffersen, K.S., S.L. Amsinck, F. Landkildehus, T.L. Lauridsen, and E. Jeppesen, 2008: Lake Flora and Fauna in Relation to Ice-Melt, Water Temperature and Chemistry at Zackenberg. In: Advances in Ecological Research. [Hans Meltofte, Torben R. Christensen, Bo Elberling,Mads C.Forchhammer and Morten Rasch (ed.)]. Academic Press, pp. 371-389. Clark, D.A., D.S. Lee, M.M.R. Freeman, and S.G. Clark, 2008: Polar Bear Conservation in Canada: Defining the Policy Problems. Arctic, 61(4), 347-360. Clarke, L.J., S.A. Robinson, Q. Hua, D.J. Ayre, and D. Fink, 2012: Radiocarbon bomb spike reveals biological effects of Antarctic climate change. Global Change Biology, 18(1), 301-310. Cochran, P., O. H. Huntington, C. Pungowiyi, S. Tom, F S. Chapin III, H. P. Huntington, N. G. Maynard, S. F. Trainor. 2013. Indigenous Frameworks for Observing and Responding to Climate Change in Alaska. Special Issue Climatic Change. Vol. 12. Issue 3. 557-567. DOI:10.1007/s10584-013-0735-2 Comeau, S., J.P. Gattuso, A.M. Nisumaa, and J. Orr, 2012: Impact of aragonite saturation state changes on migratory pteropods. Proceedings of the Royal Society B-Biological Sciences, 279(1729), 732-738. Comeau, S., R. Jeffree, J. Teyssié, and J. Gattuso, 2010: Response of the Arctic PteropodLimacina helicinato Projected Future Environmental Conditions. PLoS ONE, 5(6), e11362. Constable, A.J., 2011: Lessons from CCAMLR on the implementation of the ecosystem approach to managing fisheries. Fish and Fisheries, DOI: 10.1111/j.1467-2979.2011.00410.x. Convey, P. and M. Lebouvier, 2009: Environmental change and human impacts on terrestrial ecosystems of the sub- Antarctic islands between their discovery and the mid-twentieth century. Society, 143(1). Convey, P. and M.I. Stevens, 2007: Antarctic Biodiversity. Science, 317, 1877-1878 Convey, P., 2006: Roberto Bargagli, Ecological Studies 175: Antarctic Ecosystems Environmental Contamination, Climate Change and Human Impact. Journal of Paleolimnology, 36(2), 223-224. Convey, P., 2010: Terrestrial biodiversity in Antarctica Recent advances and future challenges. Polar Science, 4(2), 135-147. Convey, P., 2011: Antarctic terrestrial biodiversity in a changing world. Polar Biology, 34, 1629-1641. Convey, P., R. Bindschadler, G. Di Prisco, E. Fahrbach, J. Gutt, D.A. Hodgson, P.A. Mayewski, C.P. Summerhayes, J. Turner, and A. Consortium, 2009: Antarctic climate change and the environment. International Journal of Climatology, 21(6), 541-563. Convey, P., R. Key, M. Belchier, and C. Waller, 2011: Recent range expansions in non-native predatory beetles on sub-Antarctic South Georgia. Polar Biology, 34(4), 597-602. Cook, A.J., S. Poncet, A.P.R. Cooper, D.J. Herbert, and D. Christie, 2010: Glacier retreat on South Georgia and implications for the spread of rats. Antarctic Science, 22(3), 255-263. Cooper, W. Lee, Ashjian, J. Carin, Smith, L. Sharon, Codispoti, A. Louis, Grebmeier, M. Jacqueline, Campbell, G. Robert, Sherr, and B. Evelyn, 2006: Rapid Seasonal Sea-Ice Retreat in the Arctic Could Be Affecting Pacific Walrus (Odobenus rosmarus divergens) Recruitment. Aquatic Mammals, 32(1), 98-102. Corbett, J.J., D.A. Lack, J.J. Winebrake, S. Harder, J.A. Silberman, and M. Gold, 2010: Arctic shipping emissions inventories and future scenarios. Atmospheric Chemistry and Physics Discussions, 10(4), 10271-10311. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Costa, D.P., L. Huckstadt, D. Crocker, B. McDonald, M. Goebel, and M. Fedak, 2010: Approaches to studying climate change and its role on the habitat selection of Antarctic pinnipeds. Integrative and Comparative Biology, 3, 919 92 , doi:10.1093/icb/icq054. Costa, D.P., L. Huckstadt, D. Crocker, B. McDonald, M. Goebel, and M. Fedak, 2010: Approaches to studying climate change and its role on the habitat selection of Antarctic pinnipeds. Integrative and Comparative Biology, 50(6), doi:10.1093/icb/icq054. Cowan, D.A., S.L. Chown, P. Convey, M. Tuffin, K. Hughes, S. Pointing, and W.F. Vincent, 2011: Non-indigenous microorganisms in the Antarctic: assessing the risks Elsevier Trends Journals, pp. 540-548. Coyle, K.J. and L.V. Susteren, 2012: The Psychological Effects of Global Warming on the United States: Its Stresses, Traumas, and Societal Costs. National Wildlife Federation Climate Education Program, pp. 41. Coyle, K.O., L.B. Eisner, F.J. Mueter, A.I. Pinchuk, M.A. Janout, K.D. Cieciel, E.V. Farley, and A.G. Andrews, 2011: Climate change in the southeastern Bering Sea: impacts on pollock stocks and implications for the oscillating control hypothesis. Fisheries Oceanography, 20(2), 139-156. Crate, S. A. and M. Nuttall 2009: Anthropology and Climate Change: from encounters to actions, Walnut Creek, CA: Left Coast Press Crate, S.A., B.C. Forbes, L. King, and J. Kruse, 2010: Contact with nature. In: Arctic Social Indicators: a follow-up to the Arctic Human Development Report. [Larsen, J.N., P. Schweitzer, and G. Fondahl(eds.)]. TemaNord, pp. 109. Croxall, J.P., S.H.M. Butchart, B. Lascelles, A.J. Stattersfield, B. Sullivan, A. Symes, and P. Taylor, 2012: Seabird conservation status, threats and priority actions: a global assessment. Bird Conservation International, 22(1), 1. Cruikshank, J., 2001: Glaciers and Climate Change: Perspectives from Oral Tradition. Arctic, 54(4). Cubillos, J.C., S.W. Wright, G. Nash, M.F. de Salas, B. Griffiths, B. Tilbrook, A. Poisson, and G.M. Hallegraeff, 2007: Calcification morphotypes of the coccolithophorid Emiliania huxleyi in the Southern Ocean: changes in 2001 to 2006 compared to historical data. Marine Ecology-Progress Series, 348, 47-54 Cullen-Unsworth, L.C., R. Hill, J.R.A. Butler, and M. Wallace, 2011: A research process for integrating Indigenous and scientific knowledge in cultural landscapes: principles and determinants of success in the Wet Tropics World Heritage Area, Australia. The Geographical Journal, 178(4), 351-365. Curtis, T., S. Kvernmo, and P. Bjerregaard, 2005: Changing living conditions, life style and health. International Journal of Circumpolar Health, 64(5), 442. Dalpadado, P., R.B. Ingvaldsen, L.C. Stige, B. Bogstad, T. Knutsen, G. Ottersen and B. Ellertsen, 2012: Climate effects on Barents Sea ecosystem dynamics. ICES Journal of Marine Science, 69(7), 1303 1316. Danby, R.K. and D.S. Hik, 2007: Variability, contingency and rapid change in recent subarctic alpine tree line dynamics. Journal of Ecology, 95(2), 352-363. Danby, R.K., S. Koh, D.S. Hik, and L.W. Price, 2011: Four decades of plant community change in the alpine tundra of southwest Yukon, Canada. Ambio, 40(6), 660-671. Daniëls, F.J.A. and J.G. De Molenaar, 2011: Flora and vegetation of Tasiilaq, formerly Angmagssalik, southeast Greenland: A comparison of data between around 1900 and 2007. Ambio, 40(6), 650-659. Dankers, R. and H. Middelkoop, 2008: River discharge and freshwater runoff to the Barents Sea under present and future climate conditions. Climatic Change, 87(131-153). Dannevig, H., T. Rauken, and G. Hovelsrud, 2012: Implementing adaptation to climate change at the local level. Local Environment: The International Journal of Justice and Sustainability,17(6-7), 1-15. Darnis, G., D. Robert, C. Pomerleau, H. Link, P. Archambault, R.J. Nelson, M. Geoffroy, J. Tremblay, C. Lovejoy, S. H. Ferguson, B. P. V. Hunt, L. Fortier, 2012: Current state and trends in Canadian Arctic marine ecosystems: II. Heterotrophic food web, pelagic-benthic coupling, and biodiversity. Climatic Change, 115(1), 179-205. Dawson, J., E.J. Stewart, H. Lemelin, and D. Scott, 2010: The carbon cost of polar bear viewing tourism in Churchill, Canada. Journal of Sustainable Tourism, 18(3), 319-336. de Neergaard E., P. Stougaard, K. Hegh, L. Munk, 2009: Climatic changes and agriculture in Greenland: Plant diseases in potatoes and grass fields. IOP Conf. Ser.: Earth Environ. Sci. 6(37), doi:10.1088/1755- 1307/6/37/372013 Degteva, A. and C. Nellemann. 2013. Nenets migration in the landscape: impacts of industrial development in Yamal Peninsula, Russia. Pastoralism: Research, Policy and Practice 2013, 3:15. SpringerOPen Journal, http://www.pastoralismjournal.com/content/3/1/15 Derksen, C., S.L. Smith, M. Sharp, L. Brown, S. Howell, L. Copland, D.R. Mueller, Y. Gauthier, C.G. Fletcher, A. Tivy, M. Bernier, J. Bourgeois, R. Brown, C.R. Burn, C. Duguay, P. Kushner, A. Langlois, A.G. Lewkowicz, Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 A. Royer, and A. Walker, 2012: Variability and change in the Canadian cryosphere. Climatic Change, 115(1), 59-88. Derocher, A.E., M. Andersen, Wiig O, J. Aars, E. Hansen, and M. Biuw, 2011: Sea ice and polar bear den ecology at Hopen Island, Svalbard. Mar Ecol Prog Ser, 441, 273-279. Derocher, A.E., N.J. Lunn, and I. Stirling, 2004: Polar bears in a warming climate. Integrative and Comparative Biology, 44(2), 163-176. Derome, J. and N. Lukina, 2011: Interaction between environmental pollution and land-cover/land-use change in Arctic areas. In: Eurasian Arctic Land Cover and Land Use in a Changing Climate. [Gutman, G. (ed)] Springer, pp. 269-290 Déry, S.J. and E.F. Wood, 2005: Decreasing river discharge in northern Canada. Geophysical Research Letters, 32(10), - L10401. Dibike, Y., T. Prowse, T. Saloranta, and R. Ahmed, 2011: Response of Northern Hemisphere lake-ice cover and lake-water thermal structure patterns to a changing climate. Hydrological Processes, 25(19), 2942-2953. Dinniman, M.S., J.M. Klinck, and E.E. Hofmann, 2012: Sensitivity of Circumpolar Deep Water Transport and Ice Shelf Basal Melt along the West Antarctic Peninsula to Changes in the Winds. Journal of Climate, 25(14), 4799-4816. Donaldson S.G. et al, 2010. Environmental contaminants and human health in the Canadian Arctic. Science of the Total Environment 408, (2010). Donelson, J.M., P.L. Munday, M.I. Mccormick, and G.E. Nilsson, 2011: Acclimation to predicted ocean warming through developmental plasticity in a tropical reef fish. Global Change Biology, 17(4), 1712-1719. Drinkwater, K.F., 2011: The influence of climate variability and change on the ecosystems of the Barents Sea and adjacent waters: Review and synthesis of recent studies from the NESSAS project. Progress in Oceanography, 90(2011), 47-61. Drinkwater, K.F., G. Beaugrand, M. Kaeriyama, S. Kim, G. Ottersen, R.I. Perry, H. Pörtner, J.J. Polovina, and A. Takasuka, 2010: On the processes linking climate to ecosystem changes. Journal of Marine Systems, 79(3-4), 374-388. Duc, N., P. Crill, and D. Bastviken, 2010: Implications of temperature and sediment characteristics on methane formation and oxidation in lake sediments. Biogeochemistry, 100(1), 185-196. Ducklow, H., A. Clarke, R. Dickhut, S. C. Doney, H. Geisz, K. Huang, D. G. Martinson, M. P. Meredith, H. V. Moeller, M. Montes-Hugo, O. Schofield, S. E. Stammerjohn, D. Steinberg, and W. Fraser. 2012: The marine system of the western Antarctic Peninsula. Pp121-159 in [A. Rogers, N. Johnston, E. Murphy, and A. Clarke, (editors)]. Antarctic Ecosystems. Blackwell Publishing, Ltd. Duhaime, G., A. Lemelin, V. Didyk, O. Goldsmith, G. Winther, A. Caron, N. Bernard, and A. Goodmaire, 2004: Economic systems. In: The Arctic Human Development Report. Stefansson Arctic Institute, Akureyri, Iceland, pp. 69. Dumas, J.A., G.M. Flato, and R.D. Brown, 2006: Future Projections of Landfast Ice Thickness and Duration in the Canadian Arctic. Journal of Climate, 19(20), 5175-5189. Durner, G., J. Whiteman, H. Harlow, S. Amstrup, E. Regehr, and M. Ben-David, 2011: Consequences of long- distance swimming and travel over deep-water pack ice for a female polar bear during a year of extreme sea ice retreat. Polar Biology, 34(7), 975-984. Durner, G.M., D.C. Douglas, R.M. Nielson, S.C. Amstrup, T.L. McDonald, I. Stirling, M. Mauritzen, E.W. Born, O. Wiig, E. DeWeaver, M.C. Serreze, S.E. Belikov, M.M. Holland, J. Maslanik, J. Aars, D.A. Bailey, and A.E. Derocher, 2009: Predicting 21st-century polar bear habitat distribution from global climate models. Ecological Monographs, 79(1), 25-58. Dyck, M. and S. Romberg, 2007: Observations of a wild polar bear (Ursus maritimus) successfully fishing Arctic charr (Salvelinus alpinus) and Fourhorn sculpin (Myoxocephalus quadricornis). Polar Biology, 30(12), 1625- 1628. Dyck, M.G. and E. Kebreab, 2009: Estimating the Energetic Contribution of Polar Bear (Ursus Maritimus) Summer Diets to the Total Energy Budget. Journal of Mammalogy, 90(3), 585-593. Dyck, M.G., W. Soon, R.K. Baydack, D.R. Legates, S. Baliunas, T.F. Ball, and L.O. Hancock, 2007: Polar bears of western Hudson Bay and climate change: Are warming spring air temperatures the ultimate survival control factor? Ecological Complexity, 4(3), 73-84. Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Dyck, M.G., W. Soon, R.K. Baydack, D.R. Legates, S. Baliunas, T.F. Ball, and L.O. Hancock, 2008: Reply to response to Dyck et al. (2007) on polar bears and climate change in western Hudson Bay by Stirling et al. (2008). Ecological Complexity, 5(4), 289-302. Eicken, H., Lovecraft, A., Druckenmiller,M., 2009: Sea-Ice System Services: A Framework to Help Identify and Meet Information Needs Relevant for Observing Networks. Arctic, 62(2). Eicken, H., Lovecraft, A., Druckenmiller,M., 2009: Sea-Ice System Services: A Framework to Help Identify and Meet Information Needs Relevant for Observing Networks. Arctic, 62(2). Eira, I.M.G., C. Jaedicke, O.H. Magga, N. Maynard, D. Vikhamar-Schuler, and S.D. Mathiesen, 2012: Traditional Sami Snow Terminology and Physical Snow Classification Two Ways of Knowing. Cold Regions Science and Technology, Cold Regions Science and Technology 85 (117-130), doi: 10.1016/j.coldregions.2012.09.004 Ellingsen, I. H., P. Dalpadado, D. Slagstad, and H. Loeng. 2008. Impact of climatic change on the biological production in the Barents Sea. Climatic Change, 87, 155-175. Elmendorf, S.C., G.H. Henry, R.D. Hollister, R.G. Bjork, A.D. Bjorkman, T.V. Callaghan, L.S. Collier, E.J. Cooper, J.H. Cornelissen, T.A. Day, A.M. Fosaa, W.A. Gould, J. Gretarsdottir, J. Harte, L. Hermanutz, D.S. Hik, A. Hofgaard, F. Jarrad, I.S. Jonsdottir, F. Keuper, K. Klanderud, J.A. Klein, S. Koh, G. Kudo, S.I. Lang, V. Loewen, J.L. May, J. Mercado, A. Michelsen, U. Molau, I.H. Myers-Smith, S.F. Oberbauer, S. Pieper, E. Post, C. Rixen, C.H. Robinson, N.M. Schmidt, G.R. Shaver, A. Stenstrom, A. Tolvanen, O. Totland, T. Troxler, C.H. Wahren, P.J. Webber, J.M. Welker, and P.A. Wookey, 2012a: Global assessment of experimental climate Elmendorf, S.C., G.H.R. Henry, and R.D. Hollister, R.G. Björk, N. Boulanger-Lapointe, E.J. Cooper, J.H.C. Cornelissen, T.A. Day, E. Dorrepaal, T.G. Elumeeva, M. Gill, W.A. Gould, J. Harte, D.S. Hik, A. Hofgaard, D.R. Johnson, J.F. Johnstone, I.S. Jónsdóttir, J.C. Jorgenson, K. Klanderud, J.A. Klein, S. Koh, G. Kudo, M. Lara, E. Lévesque, B. Magnússon, J.L. May, J.A. Mercado-D ´az, A. Michelsen, U. Molau, I.H. Myers-Smith, S.F. Oberbauer, V.G. Onipchenko, C. Rixen, N. Martin Schmidt, G.R. Shaver, M.J. Spasojevic, .Ellen órhallsdóttir, A. Tolvanen, T. Troxler, C.E. Tweedie, S. Villareal, C-H. Wahren, X. Walker, P. J. Webber, J.M. Welker, and S. Wipf, 2012b: Plot-scale evidence of tundra vegetation change and links to recent summer warming. Nature Climate Change, 2, 453-457 Emmerton, C.A., L.F.W. Lesack, and W.F. Vincent, 2008: Mackenzie River nutrient delivery to the Arctic Ocean and effects of the Mackenzie Delta during open water conditions. Global Biogeochemical Cycles, 22(1), GB1024. Epstein, H.E., M.K. Raynolds, D.A. Walker, U.S. Bhatt, C.J. Tucker, and J.E. Pinzon, 2012: Dynamics of aboveground phytomass of the circumpolar Arctic tundra during the past three decades. Environmental Research Letters, 7(1), 015506 Epstein, P.R. and D. Ferber, 2011: Changing Planet, Changing Health: How the Climate Crisis Threatens Our Health and What We Can Do about It. University of California Press Books, pp. 368. Eskeland, G. S. and L. S. Flottorp, 2006. Climate change in the Arctic: A discussion of the impact on economic activity. In: Glomsrd, Solveig and Iulie Aslaksen (eds.), The Economy of the North. Statistics Norway. oslo, Norway. Eskeland, G. S. and L. S. Flottorp, 2006: Climate change in the Arctic: A discussion of the impact on economic activity. In: Glomsrd, Solveig and Iulie Aslaksen (eds.), The Economy of the North. Statistics Norway. Oslo, Norway. Lindholdt, L., & Glomsrd, S. (2012). The Arctic: No big bonanza for the global petroleum industry. Energy Economics, 34, 1465-1474. Fabry, V.J., J.B. McClintock, J.T. Mathis, and J.M. Grebmeier, 2009: Ocean Acidification at High Latitudes: The Bellwether. Oceanography, 22(4), 160-171. Falk-Petersen, S., P. Mayzaud, G. Kattner, and J. R. Sargent. 2009. Lipids and life strategy of Arctic Calanus. Marine Biology Research, 5:18-39. Favero-Longo, S.E., N. Cannone, M.R. Worland, P. Convey, R. Piervittori, and M. Guglielmin, 2011: Changes in lichen diversity and community structure with fur seal population increase on Signy Island, South Orkney Islands. Antarctic Science, 23, 65. Fechhelm, R.G., B. Streever, and B.J. Gallaway, 2007: The arctic cisco (Coreogonus autumnalis) subsistence and commercial fisheries, Colville River, Alaska: a conceptual model. Arctic, 60, 421-429. Ferguson, S., M. Kingsley, and J. Higdon, 2012: Killer whale (Orcinus orca) predation in a multi-prey system. Population Ecology, 54(1), 31-41. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ferguson, S.H., I. Stirling, and P. McLoughlin, 2005: Climate change and ringed seal (Phoca hispida) recruitment in western Hudson Bay. Marine Mammal Science, 21(1), 121-135. Fischbach, A.S., S.C. Amstrup, and D.C. Douglas, 2007: Landward and eastward shift of Alaskan polar bear denning associated with recent sea ice changes. Polar Biology, 30, 1395-1405. Flanagan, K.M., E. Mccauley, and F. Wrona, 2006: Freshwater food webs control carbon dioxide saturation through sedimentation. Global Change Biology, 12(4), 644-651. Flores, H., A. Atkinson, S. Kawaguchi, B.A. Krafft, G. Milinevsky, S. Nicol, C. Reiss, G.A. Tarling, R. Werner, E. Bravo Rebolledo, V. Cirelli, J. Cuzin-Roudy, S. Fielding, J.A. van Franeker, J.J. Groeneveld, M. Haraldsson, A. Lombana, E. Marschoff, B. Meyer, E.A. Pakhomov, A.P. Van de Putte, E. Rombol., K. Schmidt, V. Siegel, M. Teschke, H. Tonkes, J.Y. Toullec, P.N. Trathan, N. Tremblay, and T. Werner, 2012: Impact of climate change on Antarctic krill. Marine Ecology Progress Series, 458, 1-19. Forbes, B.C. and F. Stammler, 2009: Arctic climate change discourse: the contrasting politics of research agendas in the West and Russia. Polar Research, 28(1), 28-42. Forbes, B.C. and T. Kumpula, 2009: The ecological role and geography of reindeer (Rangifer tarandus) in Northern Eurasia. Geography Compass, 3/4, 1356-1380. Forbes, B.C., 2006: Reindeer Management in Northernmost Europe: Linking Practical And Scientific Knowledge in Social-Ecological Systems. Springer. Forbes, B.C., 2008: Equity, vulnerability and resilience in social-ecological systems: a contemporary example from the Russian Arctic. Research in Social Problems and Public Policy, 15, 203. Forbes, B.C., F. Stammler, T. Kumpula, N. Meschtyb, A. Pajunen, and E. Kaarlejärvi, 2009: High resilience in the Yamal-Nenets social-ecological system, West Siberian Arctic, Russia. Proceedings of the National Academy of Sciences of the United States of America, 106(52), 22041-22048. Forbes, B.C., M. Bölter, L. Müller-Wille, J. Hukkinen, F. Müller, N. Gunslay, Y. Konstantinov (eds.) 2006: Reindeer management in northernmost Europe: linking practical and scientific knowledge in social-ecological systems. Ecological Studies, 184. Springer-Verlag, Berlin. pp. 11-25. Forbes, B.C., M.M. Fauria, and P. Zetterberg, 2010: Russian Arctic warming and 'greening' are closely tracked by tundra shrub willows. Global Change Biology, 16(5), 1542-1554. Forbes, D.L. (ed.), 2011: State of the Arctic Coast 2010 - Scientific Review and Outlook. International Permafrost Association. HelmholtzZentrum, Geesthacht, Germany, pp. 178. Forcada, J., P.N. Trathan, P.L. Boveng, I.L. Boyd, J.M. Burns, D.P. Costa, M. Fedak, T.L. Rogers, and C.J. Southwell, 2012: Responses of Antarctic pack-ice seals to environmental change and increasing krill fishing. Biological Conservation, 149(1), 40-50. Forchhammer, M.C., N.M. Schmidt, T.T. Hye, T.B. Berg, D.K. Hendrichsen, and E. Post, 2008: Population Dynamical Responses to Climate Change. Advances in Ecological Research, 40, 391-419. Ford, J., and Goldhar, C., Climate change vulneraility and adaptation in resource dependent communities: a case study from West Greenland. Climate Research 54, 181-196. (2012). Ford, J., B. Pearce, B. Smit, J. Wandel, M. Allurut, K. Shappa, H. Ittusujurat, and K. Qrunnut, 2007: Reducing vulnerability to climate change in the Arctic: The case of Nunavut, Canada. Arctic, 60 (2), 150-166. Ford, J., B. Smit, and J. Wandel, 2006: Vulnerability to climate change in the Arctic: A case study from Arctic Bay, Nunavut. Global Environmental Change, 16 (2), 145-160. Ford, J., T. Pearce, J. Prno, F. Duerden, L. Berrang Ford, M. Beaumier, and T. Smith, 2010: Perceptions of climate change risks in primary resource use industries: a survey of the Canadian mining sector. Regional Environmental Change, 10(1), 65-81. Ford, J.D. and C. Furgal, 2009: Foreword to the special issue: climate change impacts, adaptation and vulnerability in the Arctic. Polar Research, 28(1), 1-9. Ford, J.D. and L. Berrang-Ford, 2009: Food security in Igloolik, Nunavut: an exploratory study. Polar Record, 45(03), 225. Ford, J.D. and T. Pearce, 2010: What we know, do not know, and need to know about climate change vulnerability in the western Canadian Arctic: a systematic literature review. Environmental Research Letters, 5(1), 014008. Ford, J.D., 2009: Dangerous climate change and the importance of adaptation for the Arctic's Inuit population. Environmental Research Letters, 4(2). Frederiksen, M.,Anker-Nilssen,T.,Beaugrand,G.,Wanless,S. 2013: Climate, copepods and seabirds in the boreal Northeast Atlantic - current state and future outlook. Global Change Biology, 19, 364-372 Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Freitas, C., K.M. Kovacs, R.A. Ims, and C. Lydersen, 2008: Predicting habitat use by ringed seals (Phoca hispida) in a warming Arctic. Ecological Modelling, 217(1-2), 19-32. Frenot, Y., S.L. Chown, J. Whinam, P.M. Selkirk, P. Convey, M. Skotnicki, and D.M. Bergstrom, 2005: Biological invasions in the Antarctic: extent, impacts and implications. Biological Reviews, 80(01), 45. Frey, K.E. and J.W. McClelland, 2009: Impacts of permafrost degradation on arctic river biogeochemistry. Hydrological Processes, 23(1), 169-182. Froese, R. and A. Proelß, 2010: Rebuilding fish stocks no later than 2015: will Europe meet the deadline? Fish and Fisheries, 11(2), 194-202. Fuglei, E. and R.A. Ims, 2008: Global warming and effects on the arctic fox. Science Progress, 91(2), 175-191. Furgal C., T.D. Prowse, 2008: Northern Canada. In: From Impacts to Adaptation: Canada in a Changing Climate. [Lemmen,D.S., Warren, F.J., Lacroix, J. (ed.)]. Furgal, C. and J. Seguin, 2006: Climate Change, Health, and Vulnerability in Canadian Northern Aboriginal Communities. Environ Health Perspect, 114(12). Galand, P.E., C. Lovejoy, J. Pouliot, M.E. Garneau, and W.F. Vincent, 2008: Microbial community diversity and heterotrophic production in a coastal Arctic ecosystem: A stamukhi lake and its source waters. Limnology and Oceanography, 53(2), 813. Galloway-McLean, K., 2010: Advance Guard: Climate Change Impacts, Adaptation, Mitigation and Indigenous Peoples A Compendium of Case Studies. Nations University Traditional Knowledge Initiative, Darwin, Australia, pp. 124. Gamon et al. 2013.Spatial and temporal variation in primary productivity (NDVI) of coastal Alaskan tundra: Decreased vegetation growth following earlier snowmelt. Remote Sensing of Environment,129, 144-153. GAO, 2009: U. S. General Accounting Office Report: Alaska Native Villages. Limited Progress has been Made on Relocating Villages Threatened by Flooding and Erosion. In: Report to Congressional Requesters, pp. 29-40. Gaston, A. J. 2011: Arctic seabirds: Diversity, populations, trends, and causes. in R. T. Watson, T. J. Cade, M. Fuller, G. Hunt, and E. Potapov (Eds.). Gyrfalcons and Ptarmigan in a Changing World, Volume I. The Peregrine Fund, Boise, Idaho, USA. Pp. 147-160, http://dx.doi.org/10.4080/gpcw.2011.0201 Gaston, A.J. and K. Woo, 2008: Razorbills (Alca Torda) Follow Subarctic Prey into the Canadian Arctic: Colonization Results from Climate Change? Auk, 125(4), 939-942. Gaston, A.J., H.G. Gilchrist, and J.M. Hipfner, 2005: Climate change, ice conditions and reproduction in an Arctic nesting marine bird: Brunnich's guillemot (Uria lomvia L.). Journal of Animal Ecology, 74(5), 832-841. Gaston, A.T., H.G. Gilchrist, M.L. Mallory, and P.A. Smith, 2009: Changes in Seasonal Events, Peak Food Availability, and Consequent Breeding Adjustment in a Marine Bird: a Case of Progressive Mismatching. Condor, 111(1), 111-119. Gautier, D.L., K.J. Bird, R.R. Charpentier, A. Grantz, D.W. Houseknecht, T.R. Klett, T.E. Moore, J.K. Pitman, C.J. Schenk, J.H. Schuenemeyer, K. Srensen, M.E. Tennyson, Z.C. Valin, and C.J. Wandrey, 2009: Assessment of Undiscovered Oil and Gas in the Arctic. Science, 324(5931), 1175-1179. Gearheard, S., C. Aporta, G. Aipellee, and K. O Keefe, 2011: The Igliniit project: Inuit hunters document life on the trail to map and monitor arctic change. Canadian Geographer / Le Géographe Canadien, 55(1), 42-55. Gearheard, S., G. Aipellee, and K. O'Keefe, 2010: The Igliniit project: Combining Inuit knowledge and geomatics engineering to develop a new observation tool for hunters. In: SIKU: Knowing our ice, documenting Inuit sea- ice knowledge and use. [Krupnik, I., C. Aporta, S. Gearheard, G.J. Laidler, and L. Kielsen-Holm(eds.)]. Springer, pp. 181. Gearheard, S., W. Matumeak, I. Angutikjuaq, J. Maslanik, H.P. Huntington, J. Leavitt, D.M. Kagak, G. Tigullaraq, and R.G. Barry, 2006: "It's not that simple": A collaborative comparison of sea ice environments, their uses, observed changes, and adaptations in Barrow, Alaska, USA, and Clyde River, Nunavut, Canada. Ambio, 35(4), 203-211. Geffen, A.J., H. Hie, A. Folkvord, A.K. Hufthammer, C. Andersson, U. Ninnemann, R.B. Pedersen, and K. Nedreaas, 2011: High-latitude climate variability and its effect on fisheries resources as revealed by fossil cod otoliths. ICES Journal of Marine Science: 68(6), 1081-1089. Gilg, O., B. Sittler, and I. Hanski, 2009: Climate change and cyclic predator-prey population dynamics in the high Arctic. Global Change Biology, 15(11), 2634-2652. Gilg, O., Kovacs, K.M., Aars,J., Fort,J., Gauthier,G., Grémillet,D., Ims,R.A., Meltofte,H., Moreau,J., Post,E., Schmidt,N.M., Yannic,G., and Bollache, L. 2012. Climate change and the ecology and evolution of Arctic vertebrates. Annals of the New York Academy of Sciences, 1249, 160-190 Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Gjosaeter, H., B. Bogstad,S.Tjelmeland, 2009: Ecosystem effects of the three capelin stock collapses in the Barent Sea. Marine Biology Research, 5, 40-53. Gleason, J.S. and K.D. Rode, 2009: Polar Bear Distribution and Habitat Association Reflect Long-term Changes in Fall Sea Ice Conditions in the Alaskan Beaufort Sea. Arctic, 62(4), 405-417. Glomsrd, S., and Aslaksen I. 2009. The Economy of the North 2008. Statistics Norway. Oslo, Norway Goetz, S.J., A.G. Bunn, G.J. Fiske, and R.A. Houghton, 2005: Satellite-observed photosynthetic trends across boreal North America associated with climate and fire disturbance. Proceedings of the National Academy of Sciences of the United States of America, 102(38), 13521-13525. Grant, S., A. Constable, B. Raymond, and S. Doust, 2006: Bioregionalisation of the Southern Ocean: Report of Experts Workshop, WWF- Australia and ACE CRC., Hobart, September 2006. Grebmeier, J.M., 2012: Shifting Patterns of Life in the Pacific Arctic and Sub-Arctic Seas. Annual Review of Marine Science, 4(1), 63-78. Grebmeier, J.M., J.E. Overland, S.E. Moore, E.V. Farley, E.C. Carmack, L.W. Cooper, K.E. Frey, J.H. Helle, F.A. McLaughlin, and S.L. McNutt, 2006: A Major Ecosystem Shift in the Northern Bering Sea. Science, 311(5766), 1461-1464. Green, D. and G. Raygorodetsky, 2010: Indigenous knowledge of a changing climate. Climatic Change, 100(2), 239-242. Greenwald, M.J., W.B. Bowden, M.N. Gooseff, J.P. Zarnetske, J.P. McNamara, J.H. Bradford, and T.R. Brosten, 2008: Hyporheic exchange and water chemistry of two arctic tundra streams of contrasting geomorphology. 113, G02029. Gremillet D. and Boulinier T., 2009: Spatial ecology and conservation of seabirds facing global climate change: A review. Mar.Ecol.Prog.Ser.Marine Ecology Progress Series, 391, 121-137. Grémillet, D., Welcker, J., Karnovsky, N.J., Walkusz, W.,Hall, M.E., Fort8, J., Brown, Z.W., Speakman, J.R., Harding, A.M. A. 2012. Little auks buffer impacts of current Arctic climate change. Marine Ecology Progress Series, 45, 197-206 Grenfell, T.C., and J. Putkonen (2008) A method for the detection of the severe rain-on-snow event on Banks Island, October 2003, using passive microwave remote sensing. Water Resources Research 44: W03425. Griffith, B., L.G. Adams, D.C. Douglas, C. Cuyler, R.G. White, A. Gunn, D.E. Russell, and R.D. Cameron, 2010: Science: E-Letters. In: No Evidence of Trophic Mismatch for Caribou in Greenland . Grnlund, A., 2009: Virkning av klimaendring pa arealbruk i norsk arktis. Bioforsk Rapport, 4(109). Gudasz, C., D. Bastviken, K. Steger, K. Premke, S. Sobek, and L.J. Tranvik, 2010: Temperature-controlled organic carbon mineralization in lake sediments. Nature, 466(7305), 478-481. Gunn, A., Miller, F., Barry, S., Buchan,A., 2009: A Near-Total Decline in Caribou on Prince of Wales, Somerset, and Russell Islands, Canadian Arctic. Arctic, 59(1). Gutt, J., I. Barratt, E. Domack, C.D. d'Acoz, W. Dimmler, A. Gremare, O. Heilmayer, E. Isla, D. Janussen, E. Jorgensen, K.H. Kock, L.S. Lehnert, P. Lopez-Gonzales, S. Langner, K. Linse, M.E. Manjon-Cabeza, M. Meissner, A. Montiel, M. Raes, H. Robert, A. Rose, E.S. Schepisi, T. Saucede, M. Scheidat, H.W. Schenke, J. Seiler, and C. Smith, 2011: Biodiversity change after climate-induced ice-shelf collapse in the Antarctic. Deep- Sea Research Part Ii-Topical Studies in Oceanography, 58(1-2), 74-83. Halldórsson G., B.D. Sigurdsson, O.E.S. Hrafnkelsdóttir, Ó. Eggertsson, E. Ólafsson, 2013: New anthropod herbivores on trees and shrubs in Iceland and changes in pest dynamics: A review. ICEL. AGRIC. SCI. 26 (2013), 69-84 Hallinger, M., M. Manthey, and M. Wilmking, 2010: Establishing a missing link: warm summers and winter snow cover promote shrub expansion into alpine tundra in Scandinavia. New Phytologist, 186(4), 890-899. Hansen et al. 2013. Climate events synchronize the dynamics of a resident vertebrate community in the high arctic. Science339, 313-315. Hansen, B.B., R. Aanes, I. Herfindal, J. Kohler, B.E. Saether, and M.K. Oli, 2011: Climate, icing, and wild arctic reindeer: Past relationships and future prospects. Ecology, 92(10), 1917-1923. Harsch, M.A., P.E. Hulme, M.S. McGlone, and R.P. Duncan, 2009: Are treelines advancing? A global meta-analysis of treeline response to climate warming. Ecology Letters, 12(10), 1040-1049. Harsem, O., Eide, A. & Heen, K. (2011). Factors Influencing Future Oil and Gas Prospects in the Arctic. Energy Policy. 39: 8037-8045. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hátún, H., M. R. Payne, and J. A. Jacobsen. 2009: The North Atlantic subpolar gyre regulates the spawning distribution of blue whiting (Micromesistius poutassou). Canadian Journal of Fisheries and Aquatic Sciences, 66(5):759-770. Hay, L. and G. McCabe, 2010: Hydrologic effects of climate change in the Yukon River Basin. Climatic Change, 100(3-4), 509-523. Haynie, A.C. and L. Pfeiffer, 2012: Why economics matters for understanding the effects of climate change on fisheries. ICES Journal of Marine Science, 69(7), 1160-1167. Hedenas, H., H. Olsson, C. Jonasson, J. Bergstedt, U. Dahlberg, and T.V. Callaghan, 2011: Changes in tree growth, biomass and vegetation over a 13-year period in the Swedish sub-Arctic. Ambio, 40(6), 672-682. Hedensted L., D., K. Sehested, T. Hellesen, et al., 2012. Climate change adaptation in Denmark: enhancement through collaboration and meta-governance? Local Environment. 17 (6-7), 613-28. Heggberget, T, M., Gaare E. and Ball, J.P. 2010. Reindeer (Rangifer tarandus) and climate change: Importance of winter forage. Rangifer 22, 13-31. Heino, J., R. Virkkala, and H. Toivonen, 2009: Climate change and freshwater biodiversity: detected patterns, future trends and adaptations in northern regions. Biological Reviews, 84(1), 39-54. Heliasz, M., T. Johansson, A. Lindroth, M. Mölder, M. Mastepanov, T. Friborg, T.V. Callaghan, and T.R. Christensen, 2011: Quantification of C uptake in subarctic birch forest after setback by an extreme insect outbreak. Geophysical Research Letters, 38(1), 5. Helle, T.P. and L.M. Jaakkola, 2008: Transitions in herd management of semi-domesticated reindeer in northern Finland Akateeminen kirjakauppa, Helsinki, Finlande, pp. 81; 21-101. Henden, J.-., R.A. Ims, N.G. Yoccoz, P. Hellström, and A. Angerbjörn, 2010: Strength of asymmetric competition between predators in food webs ruled by fluctuating prey: The case of foxes in tundra. Oikos, 119(1), 27-34. Hezel, P.J., X. Zhang, C.M. Bitz, B.P. Kelly, and F. Massonnet, 2012: Projected decline in spring snow depth on Arctic sea ice caused by progressively later autumn open ocean freeze-up this century. Geophysical Research Letters, 39(17), - L17505. Higdon, J.W. and S.H. Ferguson, 2009: Loss of Arctic sea ice causing punctuated change in sightings of killer whales (Orcinus orca) over the past century. Ecological Applications, 19(5), 1365-1375. Hill, G.B. and G.H.R. Henry, 2011: Responses of High Arctic wet sedge tundra to climate warming since 1980. Global Change Biology, 17(1), 276-287. Hill, S. L., T. Phillips, and A. Atkinson, 2013: Potential Climate Change Effects on the Habitat of Antarctic Krill in the Weddell Quadrant of the Southern Ocean. PloS one 8.8: e72246. Hill, Simeon L., Tony Phillips, and Angus Atkinson, 2013: Potential Climate Change Effects on the Habitat of Antarctic Krill in the Weddell Quadrant of the Southern Ocean. PloS one 8.8 (2013): e72246. Hinkel, K.M., B.M. Jones, W.R. Eisner, C.J. Cuomo, R.A. Beck, and R. Frohn, 2007: Methods to assess natural and anthropogenic thaw lake drainage on the western Arctic coastal plain of northern Alaska. 112, F02S16 Hinz, D.J., A.J. Poulton, M.C. Nielsdottir, S. Steigenberger, R.E. Korb, E.P. Achterberg, and T.S. Bibby, 2012: Comparative seasonal biogeography of mineralising nannoplankton in the Scotia Sea: Emiliania huxleyi, Fragilariopsis spp. and Tetraparma pelagica. Deep-Sea Research Part Ii-Topical Studies in Oceanography, 59, 57-66. Hinzman, LD, Bettez N, Bolton WR, Chapin III FS, Dyurgerov M, Fastie C, Griffith B, Hollister RD, Hope A, Huntington HP et al.. 2005. Evidence and implications of recent climate change in terrestrial regions of the Arctic. Climate Change. 72:251-298. Hodgson, D.A. and J.P. Smol, 2008: High latitude paleolimnology. In: Polar Lakes and Rivers - Limnology of Arctic and Antarctic Aquatic Ecosystems. [Vincent, W.F. and J. Laybourn-Parry(eds.)]. Oxford University Press, Oxford, UK, pp. 43-64; 3. Hodgson, D.A., 2011: First synchronous retreat of ice shelves marks a new phase of polar deglaciation. Proceedings of the National Academy of Sciences, 108(47), 18859-18860. Hodgson, D.A., E. Verleyen, A.H. Squier, K. Sabbe, B.J. Keely, K.M. Saunders, and W. Vyverman, 2006b: Interglacial environments of coastal east Antarctica: comparison of MIS 1 (Holocene) and MIS 5e (Last Interglacial) lake-sediment records. Quaternary Science Reviews, 25, 179-197. Hodgson, D.A., E. Verleyen, K. Sabbe, A.H. Squier, B.J. Keely, M.J. Leng, K.M. Saunders, and W. Vyverman, 2005: Late Quaternary climate-driven environmental change in the Larsemann Hills, East Antarctica, multi- proxy evidence from a lake sediment core. Quaternary Research, 64, 83-99. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hofgaard, A., J.O. Lokken, L. Dalen, and H. Hytteborn, 2010: Comparing warming and grazing effects on birch growth in an alpine environment - a 10-year experiment. Plant Ecol.Diversity Plant Ecology and Diversity, 3(1), 19-27. Hollowed, A.B., B. Planque, and H. Loeng, 2013: Potential movement of fish and shellfish stocks from the Sub- Arctic to the Arctic Ocean. Fisheries Oceanography, 22(5), 355-370. Hovelsrud, G.K. and B. Smit (eds.), 2010: Community Adaptation and Vulnerability in Arctic Regions. Springer, pp. 353. Hovelsrud, G.K., B. Poppel, B.E.H. van Oort, and J.D. Reist, 2011: Arctic Societies, Cultures, and Peoples. In: Snow, Water, Ice and Permafrost in the Arctic. AMAP, Oslo, pp. 445-483. Hye, T.T., E. Post, H. Meltofte, N.M. Schmidt, and M.C. Forchhammer, 2007: Rapid advancement of spring in the High Arctic. Current Biology, 17(12), R449-R451. Hudson, J.M. and G.H. Henry, 2009: Increased plant biomass in a High Arctic heath community from 1981 to 2008. Ecology, 90(10), 2657-63. Hueffler, K., Parkinson, A.J., Gerlach, R. and Berner, J. , 2013. Zoonotic infections in Alaska: Disease prevalence, potential impact of climate change and recommended actions for earlier disease detection, research, prevention and control. Int. J. Circumpolar Health. 72: 19562, http://dx.doi.org/10.3402/ijch.v7210.19562 Hughes, K.A. and P. Convey, 2010: The protection of Antarctic terrestrial ecosystems from inter- and intra- continental transfer of non-indigenous species by human activities: A review of current systems and practices. Global Environmental Change, 20(1), 96-112. Hughes-Hanks, J.M., L.G. Rickard, C. Panuska, J.R. Saucier, T.M. O'Hara, L. Dehn, and R.M. Rolland, 2005: Prevalence of Cryptosporidium spp. and Giardia spp. in five marine mammal species. The Journal of Parasitology, 91(5), 1225-1228. Hung, H., Kallenborn R., Breivik K., Su Y., Brorström-Lundén E., Olafsdottir K., Thorlacius J.M., Leppänen S., Bossi R., Skov H., Man S., Patton G.W., Stern G., Sverko E., Fellin P. 2010. Atmospheric monitoring of organic pollutants in the Arctic under the Arctic Monitoring and Assessment Programme (AMAP): 1993-2006. Science of the Total Environment. 408, 2854-2873. Hunt Jr., G.L., A.L. Blanchard, P. Boveng, P. Dalpadado, K.F. Drinkwater, L. Eisner, R.R. Hopcroft, K.M. Kovacs, B.L. Norcross, P. Renaud, M. Reigstad, M. Renner, H.R. Skjoldal, A. Whitehouse, and R.A. Woodgate, 2013: The Barents and Chukchi Seas: Comparison of two Arctic shelf ecosystems. Journal of Marine Systems, 109 110, 43-68. Hunt, G.L., K.O. Coyle, L.B. Eisner, E.V. Farley, R.A. Heintz, F. Mueter, J.M. Napp, J.E. Overland, P.H. Ressler, S. Salo, and P.J. Stabeno, 2011: Climate impacts on eastern Bering Sea foodwebs: a synthesis of new data and an assessment of the Oscillating Control Hypothesis. ICES Journal of Marine Science, 68(6), 1230-1243. Hunter, C.M., H. Caswell, M.C. Runge, E.V. Regehr, S.C. Amstrup, and I. Stirling, 2010: Climate change threatens polar bear populations: a stochastic demographic analysis. Ecology, 91(10), 2883-2897. Huntington, H., L. Hamilton, C. Nicolson, R. Brunner, A. Lynch, A. Ogilvie, and A. Voinov, 2007: Toward understanding the human dimensions of the rapidly changing arctic system: insights and approaches from five HARC projects. Regional Environmental Change, 7(4), 173-186. Huntington, O.H. and A. Watson, 2012: Interdisciplinarity, Native Resilience, and How the Riddles Can Teach Wildlife Law in an Era of Rapid Climate Change. Wicazo Sa Review, (2), 49. Hurst, T. P. E. R. Fernandez, J. T. Mathis, J. A. Miller, C. M. Stinson, E. F. Ahgeak, 2012: Resiliency of juvenile walleye pollock to projected levels of ocean acidification. Aquatic Biology 17: 247 259 Huse, G., I.Ellingsen, 2008: Capelin migrations and climate change - a modelling analysis. Climatic Change, 87, 177-197. IAATO, 2012: International Association of Antarctic Tour Operators - Overview of Antarctic Tourism: 2011-12 Season and Preliminary Estimates for 2012-13 Season, ATCM XXXV Information Paper 39. Ignatowski, J. and J. Rosales. 2013. Identifying the exposure of two subsistence villages in Alaska to climate hange using traditional ecological knowledge. Climatic Change. Online: DOI: 10.1007/s10584-013-0883-4. Ignatowski, J. and J. Rosales, 2012: Aksik: Stories about Adaptation and subsistence: Native voices from the Frontlines of climate change St. Lawrence University, Canton, N.Y. Ims, R.A., N.G. Yoccoz, and S.T. Killengreen, 2011: Determinants of lemming outbreaks. Proceedings of the National Academy of Sciences of the United States of America, 108(5), 1970-1974. Ims, R.A., N.G. Yoccoz, K.A. Brathen, P. Fauchald, T. Tveraa, and V. Hausner, 2007: Can reindeer overabundance cause a trophic cascade? Ecosystems, 10(4), 607-622. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ingram, K.T., et al.,: Opportunities and constraints for farmers of west Africa to use seasonal precipitation forecasts with Burkina Faso as a case study. Agricultural Systems, 74(3), (2002) IPCC, 2012: Managing the Risks of Extreme Events and Disasters to Advance ClimateChange Adaptation. A Special Report of Working Groups I and II of theIntergovernmental Panel on Climate Change [Field, C.B., V. Barros, T.F. Stocker,D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen,M. Tignor, and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, UK,and New York, NY, USA, 582 pp. Irons, D.B., T. Anker-Nilssen, A.J. Gaston, G.V. Byrd, K. Falk, G. Gilchrist, M. Hario, M. Hjernquist, Y.V. Krasnov, A. Mosbech, B. Olsen, A. Petersen, J.B. Reid, G.J. Robertson, H. Strom, and K.D. Wohl, 2008: Fluctuations in circumpolar seabird populations linked to climate oscillations. Global Change Biology, 14(7), 1455-1463. Iverson, S.J., I. Stirling, and S.L.C. Lang, 2006: Spatial and temporal variation in the diets of polar bears across the Canadian arctic: indicators of changes in prey populations and environment. In: Top Predators in Marine Ecosystems. [I. L. Boyd,S.Wanless and C.J.Camphuysen (ed.)]. Cambridge University Press, Cambridge, pp. 98. Jansson, M., A. Jonsson, A. Andersson, and J. Karlsson, 2010: Biomass and structure of planktonic communities along an air temperature gradient in subarctic Sweden. Freshwater Biology, 55(3), 691-700. Jarvis, T., N. Kelly, et al. (2010). "Acoustic characterisation of the broad-scale distribution and abundance of Antarctic krill (Euphausia superba) off East Antarctica (30-80 degrees E) in January-March 2006." Deep-Sea Research Part Ii-Topical Studies in Oceanography 57(9-10): 916-933. Jenouvrier, S., C. Barbraud, B. Cazelles, and H. Weimerskirch, 2005a: Modelling population dynamics of seabirds: importance of the effects of climate fluctuations on breeding proportions. Oikos, 108(3), 511-522. Jenouvrier, S., H. Caswell, C. Barbraud, M. Holland, J. Stroeve, and H. Weimerskirch, 2009: Demographic models and IPCC climate projections predict the decline of an emperor penguin population. Proceedings of the National Academy of Sciences of the United States of America, 106(6), 1844-1847. Jenouvrier, S., H. Weimerskirch, C. Barbraud, Y.H. Park, and B. Cazelles, 2005b: Evidence of a shift in the cyclicity of Antarctic seabird dynamics linked to climate. Proceedings of the Royal Society B-Biological Sciences, 272(1566), 887-895. Jenouvrier, S., M. Holland, J. Stroeve, C. Barbraud, H. Weimerskirch, M. Serreze, and H. Caswell, 2012: Effects of climate change on an emperor penguin population: analysis of coupled demographic and climate models. Global Change Biology, 18(9), 2756-2770. Jepsen, J.U., L. Kapari, S.B. Hagen, T. Schott, O.P.L. Vindstad, A.C. Nilssen, and R.A. Ims, 2011: Rapid northwards expansion of a forest insect pest attributed to spring phenology matching with sub-Arctic birch. Global Change Biology, 17(6), 2071-2083. Jepsen, J.U., S.B. Hagen, R.A. Ims, and N.G. Yoccoz, 2008: Climate change and outbreaks of the geometrids Operophtera brumata and Epirrita autumnata in subarctic birch forest: Evidence of a recent outbreak range expansion. Journal of Animal Ecology, 77(2), 257-264. Jernsletten, J.-. and K. Klokov, 2002: Sustainable reindeer husbandry. Arctic Council/Centre for Saami Studies, Troms, Norway. Ji, R., C.J. Ashjian, R.G. Campbell, C. Chen, G. Gao, C.S. Davis, G.W. Cowles, and R.C. Beardsley, 2012: Life history and biogeography of Calanus copepods in the Arctic Ocean: An individual-based modeling study. Progress in Oceanography, 96(1), 40-56. Johansson, C., V.A. Pohjola, C. Jonasson, and T.V. Callaghan, 2011: Multi-decadal changes in snow characteristics in sub-Arctic Sweden. Ambio, 40(6), 566-574. Joly, K., D.R. Klein, D.L. Verbyla, T.S. Rupp, and F.S. Chapin, 2011: Linkages between large-scale climate patterns and the dynamics of Arctic caribou populations. Ecography, 34(2), 345-352. Juday, G.P., 2009: Boreal Forests and Climate Change. In: Oxford Companion to Global Change. [Goudie, A. and D. Cuff(eds.)]. Oxford University Press, Oxford, UK, pp. 75-84. Kahru, M., V. Brotas, M. Manzano-Sarabia, and B.G. Mitchell, 2011: Are phytoplankton blooms occurring earlier in the Arctic? Global Change Biology, 17(4), 1733-1739. Kaplan, J.O. and M. New, 2006: Arctic climate change with a 2°C global warming: Timing, climate patterns and vegetation change. Climatic Change, 79(3-4), 213-241. Karnovsky, N., A. Harding, W. Walkusz, S. Kwasniewski, I. Goszczko, J. Wiktor, H. Routti, A. Bailey, L. McFadden, Z. Brown, G. Beaugrand, and D. Gremillet, 2010: Foraging distributions of little auks Alle alle Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 across the Greenland Sea: implications of present and future Arctic climate change. Marine Ecology-Progress Series, 415, 283-293. Kausrud, K.L., Atle Mysterud, Harald Steen, Jon Olav Vik, Eivind Ostbye, Bernard Cazelles, Erik Framstad, Anne Maria Eikeset, Ivar Mysterud, Torstein Solhy and & Nils Chr. Stenseth 2008. Linking climate change to lemming cycles. Nature 456, 93-97. Kawaguchi, S., A. Ishida, R. King, B. Raymond, N. Waller, A. Constable, S. Nicol, M. Wakita, and A. Ishimatsu. 2013. Risk maps for Antarctic krill under projected Southern Ocean acidification. Nature Climate Change. doi:10.1038/nclimate1937 Kawaguchi, S., H. Kurihara, R. King, L. Hale, T. Berli, J.P. Robinson, A. Ishida, M. Wakita, P. Virtue, S. Nicol, and A. Ishimatsu, 2011: Will krill fare well under Southern Ocean acidification? Biology Letters, 7(2), 288-291. Kawaguchi, S., S. Nicol, and A.J. Press, 2009: Direct effects of climate change on the Antarctic krill fishery. Fisheries Management and Ecology, 16(5), 424-427. Keskitalo, E., 2008: Vulnerability and adaptive capacity in forestry in northern Europe: a Swedish case study. Climatic Change, 87(1), 219-234. Keskitalo, E., 2009: Governance in vulnerability assessment: the role of globalising decision-making networks in determining local vulnerability and adaptive capacity. Mitigation and Adaptation Strategies for Global Change, 14(2), 185-201. Kharuk, V.I., K.J. Ranson, S.T. Im, and M.M. Naurzbaev, 2006: Forest-tundra larch forests and climatic trends. Russian Journal of Ecology, 37(5), 291-298. Khon, V., I. Mokhov, M. Latif, V. Semenov, and W. Park, 2010: Perspectives of Northern Sea Route and Northwest Passage in the twenty-first century. Climatic Change, 100(3), 757-768. Kilpelainen, A., H. Gregow, H. Strandman, S. Kellomaki, A. Venalainen and H. Peltola, 2010. Impacts of climate change on the risk of snow-induced forest damage in Finland. Climate Change, 99: 193-209. Kitti, H., B.C. Forbes, and J. Oksanen, 2009: Long- and short-term effects of reindeer grazing on tundra wetland vegetation. Polar Biology, 32(2), 253-261. Kokelj, S.V., B. Zajdlik, and M.S. Thompson, 2009: The impacts of thawing permafrost on the chemistry of lakes across the subarctic boreal-tundra transition, Mackenzie Delta region, Canada. Permafrost and Periglacial Processes, 20(2), 185-199. Kotwicki, S. and R.R. Lauth. 2013: Detecting temporal trends and environmentally-driven changes in the spatial distribution of groundfishes and crabs on the eastern Bering Sea shelf. Deep-Sea Research Part II: Topical Studies in Oceanography. 94, 155-175. Kovacs, K.M. and C. Lydersen, 2008: Climate change impacts on seals and whales in the North Atlantic Arctic and adjacent shelf seas. Science Progress, 91(Pt 2), 117-150. Kovacs, K.M., C. Lydersen, J.E. Overland and S.E. Moore. 2010. Impacts of changing sea-ice conditions on Arctic marine mammals. Mar. Biodivers., 41, 181-194. Krebs, C.J., 2011: Of lemmings and snowshoe hares: The ecology of northern Canada. Proceedings of the Royal Society B: Biological Sciences, 278(1705), 481-489. Kristiansen, T., K.F. Drinkwater, R.G. Lough, and S. Sundby, 2011: Recruitment variability in North Atlantic cod and match-mismatch dynamics. PLoS ONE, 6(3), e17456. Krupnik, I. and G.C. Ray, 2007: Pacific walruses, indigenous hunters, and climate change: Bridging scientific and indigenous knowledge. Deep Sea Research Part II: Topical Studies in Oceanography, 54(23-26), 2946-2957. Krupnik, I., Allison, I., Bell, R., Cutler, P., Hik, D., López-Martínez, J., Rachold, V., Sarukhanian, E., Summerhayes, C, eds. 2011: Understanding Earth s Polar Challenges:International Polar Year 2007-2008. Summary by the IPY Joint Committee. University of the Arctic publications series (4). University of the Arctic and ICSU/WMO Joint Committee for International Polar Year 2007 2008, Rovaniemi, Finland. ISBN 978-1- 896445-55-7 Krupnik, I., Aporta, C., Gearheard, S., Kielsen Holm, L., and Laidler, G. 2010. SIKU: Knowing our ice Documenting Inuit sea ice knowledge and use. New York: Springer, 501pp Krupnik, I., Ed. and D. Jolly Ed. (eds.), 2002: The Earth Is Faster Now: Indigenous Observations of Arctic Environmental Change. Frontiers in Polar Social Science. Arctic Research Consortium of the United States, 3535 College Road, Suite 101, Fairbanks, AK 99709, pp. 383. Kullman, L. and L. Öberg, 2009: Post-Little Ice Age tree line rise and climate warming in the Swedish Scandes: A landscape ecological perspective. Journal of Ecology, 97(3), 415-429. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kumpula, T., A. Pajunen, E. Kaarlejärvi, B.C. Forbes, and F. Stammler, 2011: Land use and land cover change in Arctic Russia: Ecological and social implications of industrial development. Global Environmental Change, 21(2), 550-562. Kumpula, T., B.C. Forbes, F. Stammler, and N. Meschtyb, 2012: Dynamics of a coupled system: multi-resolution remote sensing in assessing social-ecological responses during 25 years of gas field development in Arctic Russia. Remote Sensing, 4, 1046-1068. Lack, D. A. and Corbett, J. J.2012: Black carbon from ships: a review of the effects of ship speed, fuel quality and exhaust gas scrubbing, Atmos. Chem. Phys., 12, 3985-4000, doi:10.5194/acp-12-3985-2012, 2012. Laidler, G.J., T. Hirose, M. Kapfer, T. Ikummaq, E. Joamie, and P. Elee, 2011: Evaluating the Floe Edge Service: how well can SAR imagery address Inuit community concerns around sea ice change and travel safety? Canadian Geographer / Le Géographe Canadien, 55(1), 91-107. Laidre, K.L., I. Stirling, L.F. Lowry, A Wiig, M.P. Heide-JA¸rgensen, and S.H. Ferguson, 2008: quantifying the sensitivity of arctic marine mammals to climate-induced habitat change. Ecological Applications, 18, S97-S125. Landerer, F.W., J.O. Dickey, and A. Güntner, 2010: Terrestrial water budget of the Eurasian pan-Arctic from GRACE satellite measurements during 2003 2009. 115, D23115. Lantz, T.C. and S.V. Kokelj, 2008: Increasing rates of retrogressive thaw slump activity in the Mackenzie Delta region, N.W.T., Canada. Geophysical Research Letters, 35(6), L06502. Larsen, J.N. and L. Huskey, 2010: Material wellbeing in the Arctic. In: Arctic Social Indicators. [Larsen, J.N., P. Schweitzer, and G. Fondahl(eds.)]. TemaNord, pp. 47-66. Larsen, J.N., P. Schweitzer, and G. Fondahl, 2010: Arctic Social Indicators. TemaNord, 519, 160. Laurion, I., W.F. Vincent, S. MacIntyre, L. Retamal, C. Dupont, P. Francus, and R. Pienitz, 2010: Variability in greenhouse gas emissions from permafrost thaw ponds. Limonology and Oceanography, 55(1), 115. Lee, S., M. Jin, and T. Whitledge, 2010: Comparison of bottom sea-ice algal characteristics from coastal and offshore regions in the Arctic Ocean. Polar Biology, 33(10), 1331-1337. Lesack, L.F.W. and P. Marsh, 2007: Lengthening plus shortening of river-to-lake connection times in the Mackenzie River Delta respectively via two global change mechanisms along the arctic coast. Geophysical Research Letters, 34(23), L23404. Lesack, L.F.W. and P. Marsh, 2010: River-to-lake connectivities, water renewal, and aquatic habitat diversity in the Mackenzie River Delta. Water Resources Research, 46(12), W12504. Levintova, M, W.I. Zapol, N. Engmann (Eds.) (2010) Behavioral and Mental Health Research in the Arctic: A Strategy Setting Meeting. Inter J. Circumpolar Health Suppl. 5. http://www.ijch.fi/CHS/CHS_2010(5).pdf Lewis, T. and S.F. Lamoureux, 2010: Twenty-first century discharge and sediment yield predictions in a small high Arctic watershed. Global and Planetary Change, 71(1-2), 27-41. Li, W.K.W., F.A. McLaughlin, C. Lovejoy, and E.C. Carmack, 2009: Smallest algae thrive as the Arctic Ocean freshens. Science, 326(5952), 539. Lima, M., S.A. Estay, 2013: Warming effects in the western Antarctic Peninsula ecosystem: the role of population dynamic models for explaining and predicting penguin trends. Population Ecology, DOI 10.1007/s10144-013- 0386-1 Lind, S. and R.B. Ingvaldsen, 2012: Variability and impacts of Atlantic Water entering the Barents Sea from the north. Deep Sea Research Part I: Oceanographic Research Papers, 62, 70-88. Lindgren, E. and R. Gustafson, 2001: Tick-borne encephalitis in Sweden and climate change. The Lancet. 358. Issue 9275. p 16-18 Lindholt, Lars. (2006). Arctic Natural Resources in a Global Perspective. S. Glomsrd and Iulie Aslaksen (eds.), The Economy of the North, pp. 27-39. Oslo: Statistics Norway. www.ssb.no Livingston, P., K. Aydin, J.L. Boldt, A.B. Hollowed, and J.M. Napp, 2011: Alaska Marine Fisheries Management: Advancements and Linkages to Ecosystem Research. In: Ecosystem Based Management: An Evolving Perspective. [Belgrano, A. and C. Fowler(eds.)]. Cambridge University Press, Cambridge, pp. 113-152. Lloyd, A.H., A.G. Bunn, and L. Berner, 2011: A latitudinal gradient in tree growth response to climate warming in the Siberian taiga. Global Change Biology, 17(5), 1935-1945. Long, W. C., K. M. Swiney, C. Harris, H. N. Page, and R. J. Foy. 2013. Effects of ocean acidification on juvenile red king crab (Paralithodes camtschaticus) and Tanner crab (Chionoecetes bairdi) growth, condition, calcification, and survival. PLoS ONE 8:e60959. Lorenzen, E.D., D. Nogués-Bravo, L. Orlando, J. Weinstock, J. Binladen, K.A. Marske, A. Ugan, M.K. Borregaard, M.T.P. Gilbert, R. Nielsen, S.Y.W. Ho, T. Goebel, K.E. Graf, D. Byers, J.T. Stenderup, M. Rasmussen, P.F. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Campos, J.A. Leonard, K.-. Koepfli, D. Froese, G. Zazula, T.W. Stafford Jr., K. Aaris-Srensen, P. Batra, A.M. Haywood, J.S. Singarayer, P.J. Valdes, G. Boeskorov, J.A. Burns, S.P. Davydov, J. Haile, D.L. Jenkins, P. Kosintsev, T. Kuznetsova, X. Lai, L.D. Martin, H.G. McDonald, D. Mol, M. Meldgaard, K. Munch, E. Stephan, M. Sablin, R.S. Sommer, T. Sipko, E. Scott, M.A. Suchard, A. Tikhonov, R. Willerslev, R.K. Wayne, A. Cooper, M. Hofreiter, A. Sher, B. Shapiro, C. Rahbek, and E. Willerslev, 2011: Species-specific responses of Late Quaternary megafauna to climate and humans. Nature, 479(7373), 359-364. Lund, D.H., K. Sehested, T. Hellesen and V. Nellemann, 2012: Climate change adaptation in Denmark: enhancement through collaboration and meta-governance? Local Environment. 17 (6-7), 613-28, DOI:10.1080/13549839.2012.678318 Lynch, H.J., R. Naveen, P.N. Trathan, and W.F. Fagan, 2012: Spatially integrated assessment reveals widespread changes in penguin populations on the Antarctic Peninsula. Ecology, 93(6), 1367-1377. Lynn, K., J. Daigle, J. Hoffman, F. Lake, N. Michelle, D. Ranco, D. C. Viles C., G. Voggesser, P. Williams, 2013. The impacts of climate change on tribal traditional foods. Climatic Change. doi:10.1007/s10584-013-0736-1. Lyons, W.B., J. Laybourn-Parry, K.A. Welch, and J.C. Priscu, 2006: Antarctic lake systems and climate change. In: Trends in Antarctic Terrestrial and Limnetic Ecosystems. Antarctica as a Global Indicator. [Bergstrom, D.M., P. Convey, and A.H.L. Huiskes(eds.)]. Springer, Netherlands, pp. 273-295. Ma, J., H. Hung, C. Tian, and R. Kallenborn. 2001. Revolatilization of persistent organic pollutants in the Arctic induced by climate change. Nature Climate Change.. 1. 255-260. Doi:1038/nclimate1167. Ma, J., Hung, H., Tian, C and Kallenborn, R. 2011. Revolatilization of persistent organic pollutants in Maccracken, J.G., 2012: Pacific Walrus and climate change: observations and predictions. Ecology and Evolution, 2(8), 2072-2090. MacDonald, G.M., K.V. Kremenetski, and D.W. Beilman, 2008a: Climate change and the northern Russian treeline zone. Philosophical Transactions of the Royal Society B: Biological Sciences, 363(1501), 2285-2299. Macias-Fauria, M., B.C. Forbes, P. Zetterberg, and T. Kumpula, 2012: Eurasian Arctic greening reveals teleconnections and the potential for novel ecosystems. Nature Climate Change, 2, 613 618, doi:10.1038/nclimate1558. Mack, M.C., M.S. Bret-Harte, T.N. Hollingsworth, R.R. Jandt, E.A.G. Schuur, G.R. Shaver, and D.L. Verbyla, 2011: Carbon loss from an unprecedented Arctic tundra wildfire. Nature, 475(7357), 489-492. Mackey, A.P., A. Atkinson, S.L. Hill, P. Ward, N.J. Cunningham, N.M. Johnston, and E.J. Murphy, 2012: Antarctic macrozooplankton of the southwest Atlantic sector and Bellingshausen Sea: Baseline historical distributions (Discovery Investigations, 1928-1935) related to temperature and food, with projections for subsequent ocean warming. Deep-Sea Research Part Ii-Topical Studies in Oceanography, 59, 130-146. Magga, O.H., S.D. Mathiesen, R.W. Corell, and A. Oskal, 2011: Reindeer Herding, Traditional Knowledge And Adaptation To Climate Change And Loss Of Grazing Land. A project led by Norway and Association of World Reindeer Herders (WRH) in Arctic Council, Sustainable Development Working Group (SDWG). In: International Centre for Reindeer Husbandry Report 1. International Centre for Reindeer Husbandry and Association of World Reindeer Herders., Fagtrykk Idé AS, Alta, Norway. Mahoney, A., S. Gearheard, T. Oshima, and T. Qillaq, 2009: Sea Ice Thickness Measurements from a Community- Based Observing Network. Bulletin of the American Meteorological Society, 90(3), 370-377. Maldonado, J.K., C. Shearer, R. Bronen, K. Peterson and H. Lazrus. 2013. The impact of climate change on tribal communities in the US: displacement, relocation, and human rights. Climatic Change. doi: 10.1007/s10584- 013-0746-z. Mamet, S.D. and G.P. Kershaw, 2012: Subarctic and alpine tree line dynamics during the last 400 years in north- western and central Canada. Journal of Biogeography, 39(5), 855-868. Marsh, P., M. Russell, S. Pohl, H. Haywood, and C. Onclin, 2009: Changes in thaw lake drainage in the Western Canadian Arctic from 1950 to 2000. Hydrological Processes, 23(1), 145-158. Massom, R.A. and S.E. Stammerjohn, 2010: Antarctic sea ice change and variability Physical and ecological implications. Polar Science, 4(2), 149-186. Matrai, P.A., E. Olson, S. Suttles, V. Hill, L.A. Codispoti, B. Light, and M. Steele, 2013: Synthesis of primary production in the Arctic Ocean: I. Surface waters, 1954 2007. Progress in Oceanography, 110, 93 106. Matthews, C., S. Luque, S. Petersen, R. Andrews, and S. Ferguson, 2011: Satellite tracking of a killer whale (Orcinus orca) in the eastern Canadian Arctic documents ice avoidance and rapid, long-distance movement into the North Atlantic. Polar Biology, 34(7), 1091-1096. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Maynard, N.G. 2006: Satellites, settlements, and human health. In: Remote Sensing of Human Settlements (Ridd, M and J.D. Hipple, Eds) Manual of Remote Sensing. American Society of Photogrammetry and Remote Sensing. 3rd Edition, 5. Bethesda, MD. Pp. 379-399. Maynard, N.G. and G.A. Conway, 2007: A view from above: use of satellite imagery to enhance our understanding of potential impacts of climate change on human health in the Arctic. Alaska Medicine, 49(2 Suppl), 38-43. Maynard, N.G., A. Oskal, J.M. Turi, S.D. Mathiesen, I.M.G. Eira, B. Yurchak, V. Etylin, and J. Gebelein, 2011: Impacts of Arctic Climate and Land Use Changes on Reindeer Pastoralism: Indigenous Knowledge & Remote Sensing. In: Eurasian Arctic Land Cover and Land Use in a Changing Climate. [Gutman, G. (ed.)]. Springer, pp. 177-205. McLaughlin, J.B., A. DePaola, C.A. Bopp, K.A. Martinek, N.P. Napolilli, C.G. Allison, S.L. Murray, E.C. Thompson, M.M. Bird, and J.P. Middaugh, 2005: Outbreak of Vibrio parahaemolyticus Gastroenteritis Associated with Alaskan Oysters. N Engl J Med, 353(14), 1463-1470. McLeod, D., G. Hallegraeff, G. Hosie, and A. Richardson. 2012. Climate-driven range expansion of the red-tide dinoflagellate Noctiluca scintillans into the Southern Ocean. Journal of Plankton Research 34:332-337. McNamara, J.P. and D.L. Kane, 2009: The impact of a shrinking cryosphere on the form of arctic alluvial channels. Hydrological Processes, 23(1), 159-168. McNeeley, S.M., 2012: Examining barriers and opportunities for sustainable adaptation to climate change in Interior Alaska. Climatic Change, 111(3-4), 835-857. Mehlum, F. 2012: Effects of sea ice on breeding numbers and clutch size of a high arctic population of the common eider Somateria mollissima. Polar Science, 6, 143-153. Melbourne-Thomas, J., A. Constable, S. Wotherspoon, and B. Raymond, 2013: Testing paradigms of ecosystem change under climate warming in Antarctica. Plos One, 8(2), e55093. Melles, M., J. Brigham-Grette, O. Glushkova, P. Minyuk, N. Nowaczyk, and H. Hubberten, 2007: Sedimentary geochemistry of core PG1351 from Lake El gygytgyn a sensitive record of climate variability in the East Siberian Arctic during the past three glacial interglacial cycles. Journal of Paleolimnology, 37(1), 89-104. Merino, G., M. Barange, J.L. Blanchard, J. Harle, R. Holmes, I. Allen, E.H. Allison, M.C. Badjeck, N.K. Dulvy, J. Holt, S. Jennings, C. Mullon, and L.D. Rodwell, 2012: Can marine fisheries and aquaculture meet fish demand from a growing human population in a changing climate? Global Environmental Change, 22(4), 795 806. Meschtyb, N., Forbes, B., Kankaanpää,P., 2010: Social Impact Assessment along Russia's Northern Sea Route: Petroleum Transport and the Arctic Operational Platform (ARCOP). Arctic, 58(3). Mesquita, P.S., F.J. Wrona, and T.D. Prowse, 2010: Effects of retrogressive permafrost thaw slumping on sediment chemistry and submerged macrophytes in Arctic tundra lakes. Freshwater Biology, 55(11), 2347-2358. Metje, M. and P. Frenzel, 2007: Methanogenesis and methanogenic pathways in a peat from subarctic permafrost. Environmental Microbiology, 9(4), 954-964. Miller, F.L. and S.J. Barry, 2009: Long-Term Control of Peary Caribou Numbers by Unpredictable , Exceptionally Severe Snow or Ice Conditions in a Non-equilibrium Grazing System. Arctic, 62(2), 175-189. Miller, W., S.C. Schuster, A.J. Welch, A. Ratan, O.C. Bedoya-Reina, F. Zhao, H.L. Kim, R.C. Burhans, D.I. Drautz, N.E. Wittekindt, L.P. Tomsho, E. Ibarra-Laclette, L. Herrera-Estrella, E. Peacock, S. Farley, G.K. Sage, K. Rode, M. Obbard, R. Montiel, L. Bachmann, Ó Ingólfsson, J. Aars, T. Mailund, O Wiig, S.L. Talbot, and C. Lindqvist, 2012: Polar and brown bear genomes reveal ancient admixture and demographic footprints of past climate change. Proceedings of the National Academy of Sciences, 109 (36), E2382-E2390 Milliman, J.D., K.L. Farnsworth, P.D. Jones, K.H. Xu, and L.C. Smith, 2008: Climatic and anthropogenic factors affecting river discharge to the global ocean, 1951 2000. Global and Planetary Change, 62(3-4), 187-194. Milner, A.M., L.E. Brown, and D.M. Hannah, 2009: Hydroecological response of river systems to shrinking glaciers. Hydrological Processes, 23(1), 62-77. Moe, B., L. Stempniewicz, D. Jakubas, F. Angelier, O. Chastel, F. Dinessen, G.W. Gabrielsen, F. Hanssen, N.J. Karnovsky, B. Ronning, J. Welcker, K. Wojczulanis-Jakubas, and C. Bech, 2009: Climate change and phenological responses of two seabird species breeding in the high-Arctic. Marine Ecology Progress Series, 393. Moen, J., K. Aune, L. Edenius, and A. Angerbjörn, 2004: Potential effects of climate change on treeline position in the Swedish mountains. Ecology and Society, 9(1). Mokhov, I.I. and V.C. Khon, 2008: Assessment of the Northern Sea Route perspectives under climate changes on the basis of simulations with the climate models ensemble. In: Natural processes in Polar regions, Part 2. [V.M. Kotlijakov, (ed.)] IG RAS: Moscow. p. 20-27. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Molenaar, E.J., 2009: Climate Change and Arctic Fisheries. In: Climate Governance in the Arctic. [Koivurova, T., E.C.H. Keskitalo, and N. Bankes(eds.)]. Springer Netherlands, pp. 145-169. Moline, M.A., N.J. Karnovsky, Z. Brown, G.J. Divoky, T.K. Frazer, C.A. Jacoby, J.J. Torrese, and W.R. Fraser, 2008: High latitude changes in ice dynamics and their impact on polar marine ecosystems. Year in Ecology and Conservation Biology 2008, 1134, 267-319. Molnar, P.K., A.E. Derocher, T. Klanjscek, and M.A. Lewis, 2011: Predicting climate change impacts on polar bear litter size. 2, 186. Monnett, C. and J. Gleason, 2006: Observations of mortality associated with extended open-water swimming by polar bears in the Alaskan Beaufort Sea. Polar Biology, 29(8), 681-687. Montes-Hugo, M., S.C. Doney, H.W. Ducklow, W. Fraser, D. Martinson, S.E. Stammerjohn, and O. Schofield, 2009: Recent Changes in Phytoplankton Communities Associated with Rapid Regional Climate Change Along the Western Antarctic Peninsula. Science, 323(5920), 1470-1473. Moore, R.D., S.W. Fleming, B. Menounos, R. Wheate, A. Fountain, K. Stahl, K. Holm, and M. Jakob, 2009: Glacier change in western North America: influences on hydrology, geomorphic hazards and water quality. Hydrological Processes, 23(25), 42-61. Moore, S.E. and H.P. Huntington, 2008: Arctic marine mammals and climate change: impacts and resilience. Ecological Applications : A Publication of the Ecological Society of America, 18(2 Suppl), s157-165. Moore, S.E., 2008: Marine mammals as ecosystem sentinels. Journal of Mammology, 89(3), 534-540. Morán, X.A.G., Á López-Urrutia, A. Calvo-Díaz, and W.K.W. Li, 2010: Increasing importance of small phytoplankton in a warmer ocean. Global Change Biology, 16(3), 1137-1144. Moy, A.D., W.R. Howard, S.G. Bray, and T.W. Trull, 2009: Reduced calcification in modern Southern Ocean planktonic foraminifera. Nature Geoscience, 2(4), 276-280. Mueter, F.J. and M.A. Litzow, 2008: Sea ice retreat alters the biogeography of the Bering Sea continental shelf. Ecological Applications, 18(2), 309-320. Mueter, F.J., J.L. Boldt, B.A. Megrey, and R.M. Peterman, 2007: Recruitment and survival of northeast Pacific Ocean fish stocks: temporal trends, covariation, and regime shifts. Canadian Journal of Fisheries and Aquatic Sciences, 64(6), 911-927. Mueter, F.J., N.A. Bond, J.N. Ianelli, A.B. Hollowed. 2011: Expected declines in recruitment of walleye pollock (Theragra chacogramma) in the eastern Bering Sea under future climate change. ICES Journal of Marine Science, 68(6), 1284-1296. Mundy, P.R.,and D.F.Evenson, 2011: Environmental controls of phenology of high-latitude Chinook salmon populations of the Yukon River, North America, with application to fishery management. ICES Journal of Marine Science, 68(6), 1155-1164. Murphy, E. J., E. E. Hofmann, J. L. Watkins, N. M. Johnston, A. Pinones, T. Ballerini, S. L. Hill, P. N. Trathan, G. A. Tarling, R. A. Cavanagh, E. F. Young, S. E. Thorpe, and P. Fretwell. 2013: Comparison of the structure and function of Southern Ocean regional ecosystems: The Antarctic Peninsula and South Georgia. Journal of Marine Systems 109:22-42. Murphy, E., J. Watkins, P. Trathan, K. Reid, M. Meredith, S. Thorpe, N. Johnston, A. Clarke, G. Tarling, M. Collins, J. Forcada, R. Shreeve, A. Atkinson, R. Korb, M. Whitehouse, P. Ward, P. Rodhouse, P. Enderlein, A. Hirst, A. Martin, S. Hill, I. Staniland, D. Pond, D. Briggs, N. Cunningham, and A. Fleming. 2007. Spatial and temporal operation of the Scotia Sea ecosystem: a review of large-scale links in a krill centred food web. Philosophical Transactions of the Royal Society B-Biological Sciences 362:113-148. Murphy, E.J., J.L. Watkins, P.N. Trathan, K. Reid, M.P. Meredith, S.L. Hill, S.E. Thorpe, N.M. Johnston, A. Clarke, G.A. Tarling, M.A. Collins, J. Forcada, A. Atkinson, P. Ward, I.J. Staniland, D.W. Pond, R.A. Cavanagh, R.S. Shreeve, R.E. Korb, M.J. Whitehouse, P.G. Rodhouse, P. Enderlein, A.G. Hirst, A.R. Martin, D.R. Briggs, N.J. Cunningham, and A.H. Fleming, 2012b: Spatial and Temporal Operation of the Scotia Sea Ecosystem. In: Antarctic Ecosystems. [Rogers, A., N. Johnston, E. Murphy, and A. Clarke (eds.)]. Blackwell Publishing, Ltd, pp. 160-212. Murphy, E.J., R.D. Cavanagh, E.E. Hofmann, S.L. Hill, A.J. Constable, D.P. Costa, M.H. Pinkerton, N.M. Johnston, P.N. Trathan, J.M. Klinck, D.A. Wolf-Gladrow, K.L. Daly, O. Maury, and S.C. Doney, 2012a: Developing integrated models of Southern Ocean food webs: Including ecological complexity, accounting for uncertainty and the importance of scale. Progress in Oceanography, 102, 74-92. Myers-Smith, I.H., B.C. Forbes, M. Wilmking, M. Hallinger, T. Lantz, D. Blok, K.D. Tape, M. MacIas-Fauria, U. Sass-Klaassen, E. Lévesque, S. Boudreau, P. Ropars, L. Hermanutz, A. Trant, L.S. Collier, S. Weijers, J. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rozema, S.A. Rayback, N.M. Schmidt, G. Schaepman-Strub, S. Wipf, C. Rixen, C.B. Ménard, S. Venn, S. Goetz, L. Andreu-Hayles, S. Elmendorf, V. Ravolainen, J. Welker, P. Grogan, H.E. Epstein, and D.S. Hik, 2011a: Shrub expansion in tundra ecosystems: Dynamics, impacts and research priorities. Environmental Research Letters, 6(4). Myers-Smith, I.H., D.S. Hik, C. Kennedy, D. Cooley, J.F. Johnstone, A.J. Kenney, and C.J. Krebs, 2011b: Expansion of canopy-forming willows over the twentieth century on Herschel Island, Yukon Territory, Canada. Ambio, 40(6), 610-623. Nakashima, D.J., McLean, K.G., Thulstrup, H., Ameyali, R.C. and Rubis, J. 2012. Indigenous Knowledge, Marginalized Peoples and Climate Change: Foundations for Assessment and Adaptation. Technical Report prepared for the Intergovernmental Panel on Climate Change Working Group II. UNESCO/UNU: Paris/Darwin. 82 pp. Nayha, S., 2005: Environmental temperature and mortality. International Journal of Circumpolar Health, 64(5), 451-458. Nicol, S, Pauly, T, Bindoff, NL, Wright, S, Thiele, D, Hosie, GW, Strutton, PG, Woehler, E (2000a) Ocean circulation off east Antarctica affects ecosystem structure and sea-ice extent. Nature 406:504-507 Nicol, S. and A. Brierley, 2010: Through a glass less darkly new approaches for studying the distribution, abundance and biology of euphausiids. Deep-Sea Research II, 57, 496 507. Nicol, S., A. Worby, and R. Leaper, 2008: Changes in the Antarctic sea ice ecosystem: potential effects on krill and baleen whales. Marine and Freshwater Research, 59, 361-382. Nicol, S., Constable, A.J., Pauly, T. (2000b) Estimates of circumpolar abundance of Antarctic krill based on recent acoustic density measurements. CCAMLR Science 7:87-99 Nicol, S., J. Foster, and S. Kawaguchi, 2011: The fishery for Antarctic krill recent developments. Fish and Fisheries,13(1), DOI: 10.1111/j.1467-2979.2011.00406.x. Nikanorov, A., V. Bryzgalo, and G. Chernogaeva, 2007: Anthropogenically modified natural background and its formation in the Russian freshwater ecosystems. Russian Meteorology and Hydrology, 32(11), 698-710. NorAcia. 2010. Klimaendringer i Norsk Arktis Konsekvenser for Livet i Nord. Norsk Polarinstitutt Rapportserie 136 NRTEE. Canada. National Round Table on the Environment and the Economy. 2009. True North: Adapting Infrastructure to Climate Change in Northern Canada. Ottawa. NRTEE Nuttall, M., F. Berkes, B. Forbes, G. Kofinas, T. Vlassova, and G. Wenzel, 2005: Hunting, herding, fishing and gathering: indigenous peoples and renewable resource use in the Arctic. Arctic Climate Impact Assessment, (January), 649-690. Ogden, N.H., C. Bouchard, K. Kurtenbach, G. Margos, L.R. Lindsay, L. Trudel, S. Nguon, and F. Milord, 2010: Active and Passive Surveillance and Phylogenetic Analysis ofBorrelia burgdorferiElucidate the Process of Lyme Disease Risk Emergence in Canada. Environ Health Perspect, 118(7). Olech, M. and K.J. Chwedorzewska, 2011: Short Note: The first appearance and establishment of an alien vascular plant in natural habitats on the forefield of a retreating glacier in Antarctica. Antarctic Science, 23(02), 153. Olivier Gilg,Kit M. Kovacs, Jon Aars,Jerome Fort,Gilles Gauthier,David Grémillet, Rolf A. Ims,Hans Meltofte, Jérme Moreau,Eric Post,Niels Martin Schmidt, Glenn Yannic, and Lo¨ c Bollache. 2012. Climate change and the ecology and evolution of Arctic vertebrates. Annals of the New York Academy of Sciences 1249, 166-190 Olli, K., P. Wassmann, M. Reigstad, T.N. Ratkova, E. Arashkevich, A. Pasternak, P.A. Matrai, J. Knulst, L. Tranvik, R. Klais, and A. Jacobsen, 2007: The fate of production in the central Arctic Ocean top down regulation by zooplankton expatriates? Progress in Oceanography, 72(1), 84-113. Olofsson, A., O. Danell, P. Forslund, and B. Hman, 2011: Monitoring changes in lichen resources for range management purposes in reindeer husbandry. Ecological Indicators, 11(5), 1149-1159. Olofsson, J., H. Tmmervik, and T.V. Callaghan, 2012: Vole and lemming activity observed from space: vegetation cycles at a regional scale. Nature Climate Change, published online April 2012. Olofsson, J., L. Oksanen, T. Callaghan, P.E. Hulme, T. Oksanen, and O. Suominen, 2009: Herbivores inhibit climate-driven shrub expansion on the tundra. Global Change Biology, 15(11), 2681-2693. Ormseth, O.A. and B.L. Norcross, 2009: Causes and consequences of life-history variation in North American stocks of Pacific cod. ICES Journal of Marine Science, 66(2), 349-357. Orr, J.C., K. Caldeira, V. Fabry, J.P. Gattuso, P. Haugan, P. Lethodey, S. Pantoja, H.O. Pörtner, U. Riebesell, T. Trull, M. Hood, E. Urban, and W. Broadgate, 2009: Research priorities for understanding ocean acidification: Summary from the second symposium on the ocean in a high-CO2 world. Oceanography, 22(4), 182. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Oskal, A., 2008: Old livelihoods in new weather: Arctic indigenous reindeer herders face the challenges of climate change. In: Development Outreach, World Bank Institute Special Paper. World Bank, pp. 22- 25. Ostreng, W., 2006: The International Northern Sea Route Programme (INSROP): Applicable lessons learned, in Polar Record, Cambridge University Press 42(220), January, pp. 71-82 Outridge, H. Sanei, Stern, Hamilton, and F. Goodarzi, 2007: Evidence for Control of Mercury Accumulation Rates in Canadian High Arctic Lake Sediments by Variations of Aquatic Primary Productivity. Environmental Science & Technology, 41(15), 5259-5265. Overeem, I. and J.P.M. Syvitski, 2010: Shifting discharge peaks in arctic rivers, 1977-2007. Geografiska Annaler: Series A, Physical Geography, 92(2), 285-296. Overland, J.E., J.A. Francis, E. Hanna, and M. Wang. 2012. The recent shift in early summer Arctic atmospheric circulation. Geophysical Research Letters. 39. L19804, doi: 10.1029/2012g1053268. Pagano, A.M., G.M. Durner, S.C. Amstrup, K.S. Simac, and G.S. York, 2012: Long-distance swimming by polar bears (Ursus maritimus) of the southern Beaufort Sea during years of extensive open water. Canadian Journal of Zoology, 90(5), 663-676. Parada, C., D.A. Armstrong, B. Ernst, S. Hinckley, and J.M. Orensanz, 2010: Spatial dynamics of snow crab (chionoecetes opilio) in the eastern bering sea-putting together the pieces of the puzzle. Bulletin of Marine Science, 86(2), 413-437. Parkinson, A.J. and B. Evengard. 2009. Climate change, its impact on human health in the Arctic and the public health response to threats of emerging infectious diseases. Global Health Action. 2, doi:10.3402/gha.v2i0.2075. Parkinson, A.J., 2009: Sustainable Development, Climate Change and Human Health in the Arctic. In: UNESCO. Climate Change and Arctic Sustainable Development: Scientific, Social Cultural, and Educational Challenges. UNESCO, Paris, pp. 156-163. Parkinson, A.J., M.G. Bruce, and T. Zulz, 2008: International Circumpolar Surveillance, an Arctic network for the surveillance of infectious diseases. 14(1). Parnikoza, I., P. Convey, I. Dykyy, V. Trokhymets, G. Milinevsky, O. Tyschenko, D. Inozemtseva, and I. Kozeretska, 2009: Current status of the Antarctic herb tundra formation in the Central Argentine Islands. Global Change Biology, 15(7), 1685-1693. Paxian, A., V. Eyring, W. Beer, R. Sausen, and C. Wright, 2010: Present-Day and Future Global Bottom-Up Ship Emission Inventories Including Polar Routes. Environmental Science & Technology, 44(4), 1333-1339. Payette, S., 2007: Contrasted dynamics of Northern Labrador tree lines caused by climate change and migrational lag. Ecology, 88(3), 770-780. Pearson, R.G., Phillips, S.J.,, Loranty, M.M., Beck, P.S.A., Damoulas, T., Knight, S.J. and Goetz, S.J. 2013. Shifts in Arctic vegetation and associated feedbacks under climate change. Nature Climate Change 3, 673-677. Peck, L.S., D.K.A. Barnes, A.J. Cook, A.H. Fleming, and A. Clarke, 2009: Negative feedback in the cold: ice retreat produces new carbon sinks in Antarctica. Global Change Biology, 16(9), 2614-2623. Péron, C., H. Weimerskirch, and C.A. Bost, 2012: Projected poleward shift of king penguins (Aptenodytes patagonicus) foraging range at the Crozet Islands, southern Indian Ocean. Proceedings of the Royal Society B, 279, 2515-2523. Péron, C., M. Authier, C. Barbraud, K. Delord, D. Besson, and H. Weimerskirch, 2010: Interdecadal changes in at- sea distribution and abundance of subantarctic seabirds along a latitudinal gradient in the Southern Indian Ocean. Global Change Biology, 16(7), 1895-1909. Perrette, M., A. Yool, G.D. Quartly, and E.E. Popova, 2011: Near-ubiquity of ice-edge blooms in the Arctic. Biogeosciences, 8(2), 515-524. Peters G.P., Nilssen T.B., Lindholdt L., Eide m.S., Glomsr d S., Eide L.I., and Fuglestvedt J.S., 2011 : Future emissions from shipping and petroleum activities in the Arctic. Atmos. Chem. Phys., 11, 5305-5320. Peters G.P., T.B. Nilssen, L. Lindholdt, M.S. Eide, S. Glomsrd, L.I. Eide, and Fuglestvedt J.S., 2011: Future emissions from shipping and petroleum activities in the Arctic. Atmos. Chem. Phys., 11, 5305-5320. Peterson, B.J., R.M. Holmes, J.W. McClelland, C.J. Vörösmarty, R.B. Lammers, A.I. Shiklomanov, I.A. Shiklomanov, and S. Rahmstorf, 2002: Increasing River Discharge to the Arctic Ocean. Science, 298(5601), 2171-2173. Pohl, S., P. Marsh, and B. Bonsal, 2007: Modeling the Impact of Climate Change on Runoff and Annual Water Balance of an Arctic Headwater Basin. Arctic, 60(2), 173-186. Poloczanska, E.S., C.J. Brown, W.J. Sydeman, W. Kiessling, D. S. Schoeman, P.J. Moore, K. Brander, J.F. Bruno, L.B. Buckley, M.T. Burrows, C.M. Duarte, B.S. Halpern, J. Holding, C.V. Kappel, M.I. O Connor, J.M. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pandolfi, C. Parmesan, F. Schwing, S. A. Thompson, and A.J. Richardson. 2013. Global imprint of climate change on marine life. Nature Climate Change, DOI: 10.1038/NCLIMATE1958 Poppel, B. and J. Kruse, 2009: The importance of a mixed cash and harvest herding based economy to living the Arctid: an analysis based on Survey of Living Conditions int the Arctic (SLICA). In: Quality of Life in the New Milleium: Advances in the Quality-of-Life Studies. Theory and Research. Springer, Verlag, pp. 27-42. Portier, C.J., T.K. Thigpen, S.R. Carter, C.H. Dilworth, A.E. Grambsch, J. Gohlke, J. Hess, S.N. Howard, G. Luber, J.T. Lutz, T. Maslak, N. Prudent, M. Radtke, J.P. Rosenthal, T. Rowles, P.A. Sandifer, J. Scheraga, P.J. Schramm, D. Strickman, J.M. Trtanj, and P. Whung, 2010: A Human Health Perspective On Climate Change: A Report Outlining the Research Needs on the Human Health Effects of Climate Change. Research Triangle Park, NC:Environmental Health Perspectives/National Institute of Environmental Health Sciences. Pörtner, H.O. and M.A. Peck, 2010: Climate change effects on fishes and fisheries: towards a cause-and-effect understanding. Journal of Fish Biology, 77(8), 1745-1779. Post, E. and M.C. Forchhammer, 2008: Climate change reduces reproductive success of an Arctic herbivore through trophic mismatch. Philosophical Transactions of the Royal Society B: Biological Sciences, 363(1501), 2369- 2375. Post, E., C. Pedersen, C.C. Wilmers, and M.C. Forchhammer, 2008: Warming, plant phenology and the spatial dimension of trophic mismatch for large herbivores. Proceedings of the Royal Society B: Biological Sciences, 275(1646), 2005-2013. Post, E., J. Brodie, M. Hebblewhite, A.D. Anders, J.A.K. Maier, and C.C. Wilmers, 2009a: Global population dynamics and hot spots of response to climate change. Bioscience, 59(6), 489-497. Post, E., M.C. Forchhammer, M.S. Bret-Harte, T.V. Callaghan, T.R. Christensen, B. Elberling, A.D. Fox, O. Gilg, D.S. Hik, T.T. Hye, R.A. Ims, E. Jeppesen, D.R. Klein, J. Madsen, A.D. McGuire, S. Rysgaard, D.E. Schindler, I. Stirling, M.P. Tamstorf, N.J.C. Tyler, R. Van Der Wal, J. Welker, P.A. Wookey, N.M. Schmidt, and P. Aastrup, 2009b: Ecological dynamics across the arctic associated with recent climate change. Science, 325(5946), 1355-1358. Pouliot, J., P.E. Galand, C. Lovejoy and W.F. Vincent, 2009. Vertical structure of archaeal communities and the distribution of ammonia monooxygenase A gene variants in two meromictic High Arctic lakes. Environmental Microbiology. 11:687-699. Prach, K., J. Ko¹nar, J. Klime¹ová, and M. Hais, 2010: High Arctic vegetation after 70 years: A repeated analysis from Svalbard. Polar Biology, 33(5), 635-639. Prowse, T., K. Alfredsen, S. Beltaos, B. Bonsal, C. Duguay, A. Korhola, J. McNamara, W. Vincent, V. Vuglinsky, G. Weyhenmeyer, K. Walter Anthony, B. Bowden, V. Buzin, Y. Dibike, N. Gantner, L. Hinzman, L. Lia, T. Ouarda, R. Pientiz, J.D. Reist, M. Stickler, J. Weckström, and F. Wrona, 2011. Changing lake and river ice regimes: trends, effects, and implications. In: SWIPA, Snow, Water, Ice Permafrost in the Arctic. Scientific Assessment of the Arctic Monitoring and Assessment Program, Oslo, Norway, p. 6.1-6.52. Prowse, T.D. and Brown, K. 2010b. Hydro-ecological effects of changing Arctic river and lake ice covers: a review. Hydrology Research 41.6: 454-461. Prowse, T.D. and K. Brown, 2010a: Appearing and disappearing lakes in the Arctic and their impacts on biodiversity. In: Arctic Biodiversity Trends 2010 Selected indicators of change. [Kurvits, T., B. Alfthan, and E. Mork(eds.)]. CAFF International Secretariat, Akureyri, Iceland, pp. 68-70. Prowse, T.D., B.R. Bonsal, C.R. Duguay, D.O. Hessen, and V.S. Vuglinsky, 2007: River and Lake Ice. In: Global Outlook for Ice & Snow. United Nations Environment Programme, pp. 201. Prowse, T.D., C. Furgal, R. Chouinard, H. Melling, D. Milburn, and S.L. Smith, 2009: Implications of Climate Change for Economic Development in Northern Canada: Energy, Resource, and Transportation Sectors. AMBIO: A Journal of the Human Environment, 38(5), 272-281. Quayle, W. C., P. Convey, L. S. Peck, C. J. Ellis-Evans, H. G. Butler and H. J. Peat, 2013: Ecological Responses of Maritime Antarctic Lakes to Regional Climate Change. In: Antarctic Peninsula Climate Variability: Historical and Paleoenvironmental Perspectives. [E. Domack, A. Leventer, A. Burnett, R. Bindschadler, P. Convey, and M. Kirby (eds.)], American Geophysical Union: 159-170. Quesada, A. and Velazquez, D., 2012: Global Change Effects on Antarctic Lakes. In: Effects of Global Warming on Freshwater Ecosystems of the World: what can be done to reduce negative impacts? (Kumagai, M., Goldman, C.R. and Robarts, R.D. eds.). Wiley-Blackwell Ltd .pp 367-382 Quesada, A. and Velazquez, D., 2013: Global Change Effects on Antarctic Freshwater Ecosystems: The Case of Maritime Antarctic Lakes. Climatic Change and Global Warming of Inland Waters: Impacts and Mitigation for Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ecosystems and Societies, First Edition. Edited by Charles R. Goldman, Michio Kumagai and Richard D. Robarts. John Wiley & Sons. 17pp. Quesada, A., W.F. Vincent, E. Kaup, J.E. Hobbie, I. Laurion, R. Pienitz, J. López-Martínez, and J.J. Durán, 2006: Landscape control of high latitude lakes in a changing climate. In: Trends in Antarctic Terrestrial and Limnetic Ecosystems. Antarctica as a Global Indicator. [Bergstrom, D.M., P. Convey, and A.H.L. Huiskes(eds.)]. Springer, Dordrecht, pp. 221-252. Ragen, T.J., H.P. Huntington, and G.K. Hovelsrud, 2008: Conservation of arctic marine mammals faced with climate change. Ecological Applications, 18(2), S166-S174. Rawlins, M.A., H. Ye, D. Yang, A. Shiklomanov, and K.C. McDonald, 2009a: Divergence in seasonal hydrology across northern Eurasia: Emerging trends and water cycle linkages. 114, D18119. Rawlins, M.A., M.C. Serreze, R. Schroeder, X. Zhang, and K.C. McDonald, 2009b: Diagnosis of the record discharge of Arctic-draining Eurasian rivers in 2007. Environmental Research Letters, 4(4), 045011. Regehr, E.V., C.M. Hunter, H. Caswell, S.C. Amstrup, and I. Stirling, 2010: Survival and breeding of polar bears in the southern Beaufort Sea in relation to sea ice. The Journal of Animal Ecology, 79(1), 117-127. Regehr, E.V., N.J. Lunn, S.C. Amstrup, and I. Stirling, 2007: Effects of earlier sea ice breakup on survival and population size of polar bears in western Hudson Bay. Journal of Wildlife Management, 71(8), 2673-2683. Regular, P.M., G.J. Robertson, W.A. Montevecchi, F. Shuhood, T. Power, D. Ballam, and J.F. Piatt, 2010: Relative importance of human activities and climate driving common murre population trends in the Northwest Atlantic. Polar Biology, 33(9), 1215-1226. Reinert, E.S., I. Aslaksen, I.M.G. Eira, S.D. Mathiesen, H. Reindert and E.I. Turi. 2009. Adapting to climate change in Saami reindeer herding: the nation-state as problem and solution. In: W.N. Adger et al (2009): Adapting to climate change: Thresholds, Values, Governance, Pp. 417-432. Cambridge University Press. New York Reinert, H., S. Mathiesen and E. Reinert, 2008: Climate Change and Pastoral Flexibility: A Norwegian Saami Case. In: The Political Economy of Northern Regional Development Yearbook [Winther, G., G. Duhaime, J. Kruse, C. Southcott, H. Aage (eds.)], TemaNord, ISBN:978-9289320160 Reist, J.D., F.J. Wrona, T.D. Prowse, J.B. Dempson, M. Power, G. Kock, T.J. Carmichael, C.D. Sawatzky, H. Lehtonen, and R.F. Tallman, 2006: Effects of climate change and UV radiation on fisheries for Arctic freshwater and anadromous species. AMBIO: A Journal of the Human Environment, 35(7), 402-410. Rennermalm, A., E. Wood, and T. Troy, 2010: Observed changes in pan-arctic cold-season minimum monthly river discharge. Climate Dynamics, 35(6), 923-939. Revich, B. and M. Podolnaya, 2011: Thawing of permafrost may disturb historic cattle burial grounds in East Siberia. Global Health Action, 4(0). Revich, B., 2008: Climate change alters human health in Russia. Studies in Russian Economic Development. 19(3) pp. 311-317, doi:10:1134/S1075700708030106. Revich, B., N. Tokarevich, and A.J. Parkinson, 2012: Climate change and zoonotic infections in the Russian Arctic. International Journal of Circumpolar Health, 71, 18792. Revich, B.A. 2008. Climate change alters human health in Russia. Studies in Russian Economic Development. 19(3), 311-317. doi:10.1134/S1075700708030106. Revich, B.A. and D.A. Shaposhnikov, 2010: Extreme temperature episodes and mortality in Yakutsk, East Siberia. Rural and Remote Health, 10(2), 1338. Revich, B.A. and D.A. Shaposhnikov. 2012. Climate change, heat waves, and cold spells as risk factors for increased mortality in some regions of Russia. Studies on Russian Economic Development. 23(2). Pp 195-207. Doi: 10.1134/S1075700712020116 Rice, J.C. and S.M. Garcia, 2011: Fisheries, food security, climate change, and biodiversity: characteristics of the sector and perspectives on emerging issues. ICES Journal of Marine Science,68(6), 1343-1353. Roberts, D., W. R. Howard, A. D. Moy, J. L. Roberts, T. W. Trull, S. G. Bray, and R. R. Hopcroft. 2011. Interannual pteropod variability in sediment traps deployed above and below the aragonite saturation horizon in the Sub-Antarctic Southern Ocean. Polar Biology 34:1739-1750. Robertson, C., T.A. Nelson, D.E. Jelinski, M.A. Wulder, and B. Boots, 2009: Spatial-temporal analysis of species range expansion: The case of the mountain pine beetle, Dendroctonus ponderosae. Journal of Biogeography, 36(8), 1446-1458. Robinson, R., T. Smith, B.J. Kirschhoffer, and C. Rosa, 2011: Polar bear (Ursus maritimus) cub mortality at a den site in northern Alaska. Polar Biology, 34(1), 139-142. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rockwell, R.F. and L.J. Gormezano, 2009: The early bear gets the goose: climate change, polar bears and lesser snow geese in western Hudson Bay. Polar Biology, 32(4), 539-547. Rode, K., E. Peacock, M. Taylor, I. Stirling, E. Born, K. Laidre, and O Wiig, 2012: A tale of two polar bear populations: ice habitat, harvest, and body condition. Population Ecology, 54(1), 3-18. Rode, K.D., J.D. Reist, E. Peacock, and I. Stirling, 2010b: Comments in response to "Estimating the energetic contribution of polar bear (Ursus maritimus) summer diets to the total energy budget" by Dyck and Kebreab (2009). Journal of Mammalogy, 91(6), 1517-1523. Rode, K.D., S.C. Amstrup, and E.V. Regehr, 2010a: Reduced body size and cub recruitment in polar bears associated with sea ice decline. Ecological Applications, 20(3), 768-782. Rogers, A., N. Johnston, E. Murphy, and A. Clarke (eds.), 2012: Antarctic Ecosystems: An Extreme Environment in a Changing World. John Wiley & Sons, Ltd. Roturier, S. and M. Roué, 2009: Of forest, snow and lichen: Sámi reindeer herders knowledge of winter pastures in northern Sweden. Forest Ecology and Management, 258(9), 1960-1967. Royles, J., M.J., A., Convey, P., Griffiths, H., Hodgson, D.A., Leng, M.J. and Charman, D.J., 2013. Plants and Soil Microbes Respond to Recent Warming on the Antarctic Peninsula. Current Biology, 23(17), http://dx.doi.org/10.1016/j.cub.2013.07.011. Rudberg, P.M., O. Wallgren and A.G. Swartling, 2012. Beyond generic adaptive capacity: exploring the adaptation space of the water supply and wastewater sector of the Stockholm region, Sweden. Climatic Change. 114 (3-4), 707-21. Rundqvist, S., H. Hedenas, A. Sandström, U. Emanuelsson, H. Eriksson, C. Jonasson, and T.V. Callaghan, 2011: Tree and shrub expansion over the past 34 years at the tree-line near Abisko, Sweden. Ambio, 40(6), 683-692. Russell, D. and A. Gunn, 2010: Reindeer herding. In: Arctic Biodiversity Trends 2010 Selected indicators of change. [Kurvits, T., B. Alfthan, and E. Mork(eds.)]. CAFF International Secretariat, Akureyri, Iceland. Rybraten, S. and G. Hovelsrud, 2010: Local Effects of Global Climate Change: Differential Experiences of Sheep Farmers and Reindeer Herders in Unjárga/Nesseby, a Coastal Sámi Community in Northern Norway. In: Community Adaptation and Vulnerability in Arctic Regions. [Hovelsrud, G.K. and B. Smit(eds.)]. Springer Netherlands, pp. 313-333. Rylander, R. and R.S.F. Schilling, 2011: Diseases caused by organic dusts. In: 10. Respiratory System. Encyclopedia of occupational health and Safety. J.M. Stellman, Editor in Chief. International Labor Organization Geneva. 7 pp. Saba, G. K., O. Schofield, J. J. Torres, E. H. Ombres, and D. K. Steinberg. 2012: Increased feeding and nutrient excretion of adult Antarctic krill, Euphausia superba, exposed to enhanced carbon dioxide (CO2). PLoS ONE 7:e52224. Sahanatien, V. and A.E. Derocher, 2012: Monitoring sea ice habitat fragmentation for polar bear conservation. Animal Conservation, 15: 397 406. doi: 10.1111/j.1469-1795.2012.00529.x Salick, J. and N. Ross, 2009: Traditional peoples and climate change. Global Environmental Change, 19(2), 137- 139. Salonen, J.S., H. Seppä, M. Väliranta, V.J. Jones, A. Self, M. Heikkilä, S. Kultti, and H. Yang, 2011: The Holocene thermal maximum and late-Holocene cooling in the tundra of NE European Russia. Quaternary Research, 75(3), 501-511. Schliebe, S., K.D. Rode, J.S. Gleason, J. Wilder, K. Proffitt, T.J. Evans, and S. Miller, 2008: Effects of sea ice extent and food availability on spatial and temporal distribution of polar bears during the fall open-water period in the Southern Beaufort Sea. Polar Biology, 31(8), 999-1010. Schmidt, K., A. Atkinson, S. Steigenberger, S. Fielding, M. Lindsay, D. Pond, G. Tarling, T. Klevjer, C. Allen, S. Nicol, and E. Achterberg, 2011: Seabed foraging by Antarctic krill: Implications for stock assessment, bentho- pelagic coupling, and the vertical transfer of iron. Limnology and Oceanography, 56(4), 1411 1428. Schneider, P. and S.J. Hook, 2010: Space observations of inland water bodies show rapid surface warming since 1985. Geophysical Research Letters, 37(22), L22405. Shaposhnikov, D., B. Revich, V. Meleshko, V. Govorkova, and T. Pavlova, 2010: Climate Change May Reduce Annual Temperature-Dependent Mortality in Subarctic: A Case Study of Archangelsk, Russian Federation. Environment and Natural Resources Research, 1(1), 75. Shearer, C. 2012. The political ecology of climate adaptation assistance: Alaska Natives, displacement, and relocation. Journal of Political Ecology. 19. 174-183. [http://jpe.library.arizona.edu/volume_19/Shearer.pdf] Shearer, C., 2011: Kivalina: a climate change story. Haymarket Books, Chicago. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sherman, K., I.M. Belkin, K.D. Friedland, J. O Reilly and K. Hyde, 2009. Accelerated warming and emergent trends in fsheries biomass yields of the world s large marine ecosystems. Ambio, 38:215-224. Sherman, K., I.M. Belkin, K.D. Friedland, J. O Reilly and K. Hyde, 2009: Accelerated warming and emergent trends in fsheries biomass yields of the world s large marine ecosystems. Ambio, 38:215-224. Shiklomanov, A.I., R.B. Lammers, M.A. Rawlins, L.C. Smith, and T.M. Pavelsky, 2007: Temporal and spatial variations in maximum river discharge from a new Russian data set. 112, G04S53. Shiyatov, S.G., M.M. Terent'ev, V.V. Fomin, and N.E. Zimmermann, 2007: Altitudinal and horizontal shifts of the upper boundaries of open and closed forests in the Polar Urals in the 20th century. Russian Journal of Ecology, 38(4), 223-227. Shreeve, R.S., M.A. Collins, G.A. Tarling, C.E. Main, P. Ward, and N.M. Johnston, 2009: Feeding ecology of myctophid fishes in the northern Scotia Sea. Marine Ecology-Progress Series, 386, 221-236. Sigler, M.F., M. Renner, S.L. Danielson, L.B. Eisner, R.R. Lauth, K.J. Kuletz, E.A. Logerwell, and Hunt George L. Jr., 2011: Fluxes, fins, and feathers: Relationships among the Bering, Chukchi, and Beaufort Seas in a time of climate change. Oceanography, 24(3), 250-265. Sigurdsson, B.D., A. Snorrason, B. Kjartansson and J.A. Jonsson, 2007: Total area of planted forests in Iceland and their carbon stocks and fluxes. In: Effects of afforestation on ecosystems, land-scape and rural development. [Halldórsson G., Oddsdóttir E.S. and Eggertsson O. (eds.)], TemaNord 508, pp. 211-217. Simpson, S., S. Jennings, M. Johnson, J. Blanchard, P. Sch n, D. Sims, and M. Genner, 2011: Continental Shelf- Wide Response of a Fish Assemblage to Rapid Warming of the Sea. Cell Press, pp. 1565-1570. Siniff, D. B., R. A. Garrott, J. J. Rotella, W. R. Fraser, and D. G. Ainley. 2008: Opinion: Projecting the effects of environmental change on Antarctic seals. Antarctic Science 20:425-435. Slagstad, D., I.H. Ellingsen, and P. Wassmann, 2011: Evaluating primary and secondary production in an Arctic Ocean void of summer sea ice: An experimental simulation approach. Progress in Oceanography, 90(1 4), 117-131. Slater, G.J., B. Figueirido, L. Louis, P. Yang, and B. van Valkenburgh, 2010: Biomechanical Consequences of Rapid Evolution in the Polar Bear Lineage. PLoS ONE, 5(11), e13870, 1-7. Smith Jr, W., D. Ainley, R. Cattaneo-Vietti, and E. Hofmann, 2012: The Ross Sea Continental Shelf: regional biogeochemical cycles, trophic interactions, and potential future changes. In: Antarctic ecosystems: an extreme environment in a changing world. [Rogers, A., N. Johnston, E. Murphy, and A. Clarke(eds.)]. Blackwell Publishing, London, pp. 213-242; 7. Smith, J.A., D.A. Hodgson, M.J. Bentley, E. Verleyen, M.J. Leng, and S.J. Roberts, 2006: Limnology of two Antarctic epishelf lakes and their potential to record periods of ice shelf loss. Journal of Paleolimnology, 35, 373 394. Smith, Jr. W., D. Ainley, R. Cattaneo-Vietti, and E. Hofmann, 2012a: The Ross Sea Continental Shelf: regional biogeochemical cycles, trophic interactions, and potential future changes. In: Antarctic ecosystems: an extreme environment in a changing world. [Rogers, A., N. Johnston, E. Murphy, and A. Clarke(eds.)]. Blackwell Publishing, London, pp. 213-242; 7. Smith, L.C., T.M. Pavelsky, G.M. MacDonald, A.I. Shiklomanov, and R.B. Lammers, 2007: Rising minimum daily flows in northern Eurasian rivers: A growing influence of groundwater in the high-latitude hydrologic cycle. 112, G04S47. Smith, P.A., K.H. Elliott, A.J. Gaston, and H.G. Gilchrist, 2010: Has early ice clearance increased predation on breeding birds by polar bears? Polar Biology, 33(8), 1149-1153. Smith, S.L. and D.W. Riseborough, 2010: Modelling the Thermal Response of Permafrost Terrain to right-of-way disturbance and climate warming. Cold Regions Science and Technology, 60(1), 92-103 Smol, J.P. and M.S.V. Douglas, 2007a: Crossing the final ecological threshold in high Arctic ponds. Proceedings of the National Academy of Sciences of the United States of America, 104(30), 12395-12397. Smol, J.P. and M.S.V. Douglas, 2007b: From controversy to consensus: making the case for recent climate change in the Arctic using lake sediments. Frontiers in Ecology and the Environment, 5(9), 466-474, doi:10.1890/060162. Solomon, S.D., D. Qin, M. Manning, Z. Chen, M. MArquie, K.B. Averyt, M. Tignor, and H.L. Miller (eds.), 2007: Climate change 2007: The physical science basis. Contribution of working group I to the fourth assessment report of the intergovernmental panel on climate change. Cambridge University Press, Cambridge, United Kingdom, pp. 996. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sreide, J.E., E. Leu, J. Berge, M. Graeve, and S. Falk-Petersen, 2010: Timing of blooms, algal food quality and Calanus glacialis reproduction and growth in a changing Arctic. Global Change Biology, 16(11), 3154-3163. St. Jacques, J. and D.J. Sauchyn, 2009: Increasing winter baseflow and mean annual streamflow from possible permafrost thawing in the Northwest Territories, Canada. Geophysical Research Letters, 36(1), L01401. Stabeno, P.J., E.V. Farley Jr., N.B. Kachel, S. Moore, C.W. Mordy, J.M. Napp, J.E. Overland, A.I. Pinchuk, and M.F. Sigler., 2012b: Comparison of the physics of the northern and southern shelves of the eastern Bering Sea and some implications for the ecosystem. Deep Sea Research Part II: Topical Studies in Oceanography, 65 70, 14-30. Stabeno, P.J., N.B. Kachel, S.E. Moore, J.M. Napp, M. Sigler, A. Yamaguchi, and A.N. Zerbini, 2012a: Comparison of warm and cold years on the southeastern Bering Sea shelf and some implications for the ecosystem. Deep Sea Research Part II: Topical Studies in Oceanography, 65-70, 31-45. Stammler, F., 2005: Reindeer Nomads Meet the Market: Culture, Property and Globalisation at the End of the Land. ISBN-13: 978-3825880460 Stephenson, S.R., L.C. Smith, and J.A. Agnew, 2011: Divergent long-term trajectories of human access to the Arctic. 1(3), 156-160. Stewart E.J., S.E.L. Howell, D. Draper, J. Yackel And A. Tivy (2007): Sea Ice in Canada s Arctic: Implications for Cruise Tourism. ARCTIC VOL. 60(4), P. 370 380 Stewart, E. J., Howell, S. E. L, Dawson, J. D., Tivy, A. & Draper, D. (2010) Cruise tourism and sea ice in Canada's Hudson Bay region. ARCTIC. (63) 57-66. Stien, A., L.E. Loel, A. Mysterud, T. Severinsen, J. Kohler, and R. Langvatn, 2010: Icing events trigger range displacement in a high-arctic ungulate. Ecology, 91(3), 915-920. Stien, A., Rolf A. Ims, Steve D. Albon, Eva Fuglei, R. Justin Irvine, Erik Ropstad, Odd Halvorsen, Rolf Langvatn, Leif Egil Loe, Vebjrn Veiberg, Nigel G. Yoccoz . 2012. Congruent responses to weather variability in high Arctic herbivores. Stirling, I. and C.L. Parkinson, 2006: Possible Effects of Climate Warming on Selected Populations of Polar Bears (Ursus maritimus) in the Canadian Arctic. Arctic, 59(3), 261-275. Stirling, I., A.E. Derocher, W.A. Gough, and K. Rode, 2008b: Response to Dyck et al. (2007) on polar bears and climate change in western Hudson Bay. Ecological Complexity, 5(3), 193-201. Stirling, I., E. Richardson, G.W. Thiemann, and A.E. Derocher, 2008a: Unusual predation attempts of polar bears on ringed seals in the southern Beaufort Sea: Possible significance of changing spring ice conditions. Arctic, 61(1), 14-22. Stirling, I., T.L. McDonald, E.S. Richardson, E.V. Regehr, and S.C. Amstrup, 2011: Polar bear population status in the northern Beaufort Sea, Canada, 1971-2006. Ecological Applications : A Publication of the Ecological Society of America JID - 9889808, (3). Stram, D.L. and D.C.K. Evans, 2009: Fishery management responses to climate change in the North Pacific. ICES Journal of Marine Science: Journal Du Conseil, 66(7), 1633-1639. Stroeve, J., M.M. Holland, W. Meier, T. Scambos, and M. Serreze, 2007: Arctic sea ice decline: Faster than forecast. Geophysical Research Letters, 34. Sundby, S. and O. Nakken, 2008: Spatial shifts in spawning habitats of Arcto-Norwegian cod related to multidecadal climate oscillations and climate change. ICES Journal of Marine Science, 65(6), 953-962. Swann, A. L., Fung, I. Y., Levis, S., Bonan, G., and Doney, S.2010. Changes in Arctic vegetation induce high- latitude warming through the greenhouse effect. Proceedings of the National Academy of Sciences, 107, 1295 1300, doi:10.1073/pnas.0913846107. Sydneysmith, R., M. . Andrachuk, B. Smit, and G.K. Hovelsrud, 2010: Vulnerability and Adaptive Capacity in Arctic Communities. In: Adaptive Capacity and Environmental Governance. [Armitage, D. and R. Plummer(eds.)]. Springer, Berlin, pp. 133-156. Tan, A., J.C. Adam, and D.P. Lettenmaier, 2011: Change in spring snowmelt timing in Eurasian Arctic rivers. 116, D03101 Tanabe, Y., S. Kudoh, S. Imura, and M. Fukuchi, 2007: Phytoplankton blooms under dim and cold conditions in freshwater lakes of East Antarctica. Polar Biology, 31(2), 199-208. Tchebakova, N.M. , G.E. Rehfeldt, and E.I. Parfenova, 2009: From Vegetation Zones to Climatypes: Effects of Climate Warming on Siberian Ecosystems. In: Permafrost Ecosystems: Siberian Larch Forest, ISBN:978-1- 4020-9692-1 Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Teng, Y., Z. Xu, Y. Luo, and F. Reverchon, 2012: How do persistent organic pollutants be coupled with biogeochemical cycles of carbon and nutrients in terrestrial ecosystems under global climate change? Journal of Soils and Sediments, 12(3), 411-419. Teran, T., L. Lamon, and A. Marcomini. 2012: Climate change effects on POPs environmental behavior: a scientific perspective for future regulatory actions. Atmospheric Pollution Research. 3. 466-476. Terauds, A., S.L. Chown, F. Morgan, H. J. Peat, D.J. Watts, H. Keys, P. Convey, and D.M. Bergstrom, 2012: Conservation biogeography of the Antarctic. Diversity and Distributions, 18(7), 726-741. the Arctic induced by climate change. Nature Climate Change 1, 255-260. Thiemann, G.W., S.J. Iverson, and I. Stirling, 2008: Polar Bear Diets and Arctic Marine Food Webs: Insights from Fatty Acid Analysis. Ecological Monographs, 78(4), 591-613. Tholstrup, L. and R.O. Rasmussen, 2009. Climate Change - and the North Atlantic. 130 pp. Norden, Stockholm, Sweden. Thompson, M., Wrona, F., Prowse,T., 2012: Shifts in Plankton, Nutrient and Light Relationships in Small Tundra Lakes Caused by Localized Permafrost Thaw. Arctic, 65(4). Thompson, M.S., S.V. Kokelj, T.D. Prowse and F.J. Wrona, 2008. The impact of sediments derived from thawing permafrost on tundra lake water chemistry: An experimental approach. In D.L. Kane and K.M. Hinkel (eds.), Proceedings of the Ninth International Conference on Permafrost, Fairbanks, Alaska. p. 1763-1768. Thrush, S.F. and V.J. Cummings, 2011: Massive icebergs, alteration in primary food resources and change in benthic communities at Cape Evans, Antarctica. Marine Ecology-an Evolutionary Perspective, 32(3), 289-299. Tokarevich, N., A. Tronin, O. Blinova, R. Buzinov, V. Boltenkov, E. Yurasova, and J. Nurse, 2011: The impact of climate change on the expansion of Ixodes persulcatus habitat and the incidence of tickborne encephalitis in the north of European Russia. Global Health Action, 4. Tmmervik, H., B. Johansen, J.A. Riseth, S.R. Karlsen, B. Solberg, and K.A. Hgda, 2009: Above ground biomass changes in the mountain birch forests and mountain heaths of Finnmarksvidda, northern Norway, in the period 1957-2006. Forest Ecology and Management, 257(1), 244-257. Tmmervik, H., J.W. Bjerke, E. Gaare, B. Johansen, and D. Thannheiser, 2012: Rapid recovery of recently overexploited winter grazing pastures for reindeer in northern Norway. Fungal Ecology, 5(1), 3-15. Towns, L., A.E. Derocher, I. Stirling, and N.J. Lunn, 2010: Changes in Land Distribution of Polar Bears in Western Hudson Bay. Arctic, 63(2), 206-212. Towns, L., A.E. Derocher, I. Stirling, N.J. Lunn, and D. Hedman, 2009: Spatial and temporal patterns of problem polar bears in Churchill, Manitoba. Polar Biology, 32(10), 1529-1537. Trathan, P. and K. Reid (2009). "Exploitation of the marine ecosystem in the sub-Antarctic: historical impacts and current consequences." Papers and Proceeding of the Royal Society of Tasmania 143: 9-14. Trathan, P., P. Fretwell, and B. Stonehouse, 2011: First recorded loss of an emperor penguin colony in the recent period of Antarctic regional warming: implications for other colonies. PLoS ONE, 6, e14738. Trathan, P.N. and D. Agnew, 2010: Climate change and the Antarctic marine ecosystem: an essay on management implications. Antarctic Science, 22(4), 387-398. Trathan, P.N., J. Forcada, and E.J. Murphy, 2007: Environmental forcing and Southern Ocean marine predator populations: effects of climate change and variability. Philosophical Transactions of the Royal Society B- Biological Sciences, 362, 2351-2365. Trathan, P.N., J. Forcada, and E.J. Murphy, 2012: Environmental forcing and Southern Ocean marine predator populations. In: Antarctic Ecosystems. [Rogers, A., N. Johnston, E. Murphy, and A. Clarke(eds.)]. Blackwell Publishing, Ltd, pp. 335-354. Tremblay, J., D. Robert, D.E. Varela, C. Lovejoy, G. Darnis, R.J. Nelson, and A.R. Sastri, 2012: Current state and trends in Canadian Arctic marine ecosystems: I. Primary production. Climatic Change, 115(1), 161-178. Trivelpiece, W.Z., J.T. Hinke, A.K. Miller, C.S. Reiss, S.G. Trivelpiece, and G.M. Watters, 2011: Variability in krill biomass links harvesting and climate warming to penguin population changes in Antarctica. Proceedings of the National Academy of Sciences of the United States of America, 108(18), 7625-7628. Turner, J. A., N. E. Barrand, T. J. Bracegirdle, P. Convey, D. A. Hodgson, M. Jarvis, A. Jenkins, G. Marshall, M. P. Meredith, H. Roscos, J. Shanklin, J. French, H. Goosse, M. Guglielmin, J. Gutt, S. Jacobs, M. C. Kennicutt, V. Masson-Delmotte, P. Mayewski, F. Navarro, S. Robinson, T. Scambos, M. Sparrow, C. Summerhayes, K. Speer, and A. Klepnikov. 2013. Antarctic climate change and the environment: An update. Polar Record:1-23. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Turner, J., R.A. Bindschadler, P. Convey, G. Di Prisco, E. Fahrbach, J. Gutt, D.A. Hodgson, P.A. Mayewski, and C.P. Summerhayes (eds.), 2009: Antarctic Climate Change and the Environment. Scientific Committee on Antarctic Research, ISBN 978-0-948277-22-1, Cambridge, pp. 526. Turunen, M., P. Soppela, H. Kinnunen, M.-. Sutinen, and F. Martz, 2009: Does climate change influence the availability and quality of reindeer forage plants? Polar Biology, 32(6), 813-832. Tyler, N.J.C., 2010: Climate, snow, ice, crashes, and declines in populations of reindeer and caribou (Rangifer tarandus L.). Ecological Monographs, 80(2), 197-219. Tyler, N.J.C., J.M. Turi, M.A. Sundset, K.S. Bull, M.N. Sara, E. Reinert, N. Oskal, C. Nellemann, J.J. McCarthy, S.D. Mathiesen, M.L. Martello, O.H. Magga, G.K. Hovelsrud, I. Hanssen-Bauer, N.I. Eira, I.M.G. Eira, and R.W. Corell, 2007: Saami reindeer pastoralism under climate change: Applying a generalized framework for vulnerability studies to a sub-arctic social-ecological system. Global Environmental Change-Human and Policy Dimensions, 17(2), 191-206. Tyler, N.J.C., M.C. Forchhammer, and N.A. Oritsland, 2008: Nonlinear effects of climate and density in the dynamics of a fluctuating population of reindeer. Ecology, 89(6), 1675-1686. UNEP/AMAP, 2011: Climate Change and POPs: Predicting the Impacts. Report of the UNEP/AMAP Expert Group. Secretariat of the Stockholm Convention. Geneva 62 pp. USARC. 2010. Report on Goals and Objectives for Arctic Research 2009-2010. US Arctic Research Commission. Arlington, VA. 52 pp. Valdimarsson, H., O.S. Astthorsson, and J. Palsson, 2012: Hydrographic variability in Icelandic waters during recent decades and related changes in distribution of some fish species. ICES Journal of Marine Science, 69(5), 815- 825. Valdivia C, Seth et.al.: Adapting to Climate Change in Andean Ecosystems: Landscapes, Capitals, and Perceptions Shaping Rural Livelihood Strategies and Linking Knowledge Systems. Annals of the Association of American Geographers, 100(4), (2010). Van Bogaert, R., C. Jonasson, M. De Dapper, and T.V. Callaghan, 2010: Range expansion of thermophilic aspen (Populus tremula L.) in the Swedish subarctic. Arctic, Antarctic, and Alpine Research, 42(3), 362-375. Van Bogaert, R., K. Haneca, J. Hoogesteger, C. Jonasson, M. De Dapper, and T.V. Callaghan, 2011: A century of tree line changes in sub-Arctic Sweden shows local and regional variability and only a minor influence of 20th century climate warming. Journal of Biogeography, 38(5), 907-921. van Der Wal, R., S. Sjögersten, S.J. Woodin, E.J. Cooper, I.S. Jónsdóttir, D. Kuijper, T.A.D. Fox, and A.D. Huiskes, 2007: Spring feeding by pink-footed geese reduces carbon stocks and sink strength in tundra ecosystems. Global Change Biology, 13(2), 539-545. Veillette, J., D.R. Mueller, D. Antoniades, and W.F. Vincent, 2008: Arctic epishelf lakes as sentinel ecosystems: Past, present and future. 113, G04014. Velázquez, D., A. Frias, M.A. Lezcano and A. Quesada, 2013: Ecological relationships and stoichiometry within a maritime Antarctic watershed. Antarctic Science 25(2), 191-197 Verbyla, D., 2008: The greening and browning of Alaska based on 1982-2003 satellite data. Global Ecology and Biogeography, 17(4), 547-555. Vert-pre, K.A., R.O. Amoroso, O.P. Jensen, and R. Hilborn. 2013. Frequency and intensity of productivity regime shifts in marine stocks. Proceedings of the National Academy of Sciences, 110(5), 1779-1784. Vieira, G., J. Bockheim, M. Guglielmin, M. Balks, A.A. Abramov, J. Boelhouwers, N. Cannone, L. Ganzert, D.A. Gilichinsky, S. Goryachkin, J. Lopez-Martinez, I. Meiklejohn, R. Raffi, M. Ramos, C. Schaefer, E. Serrano, F. Simas, R. Sletten, D. Wagner, 2010: Thermal State of permafrost and active-layer monitoring in the Antarctic: advances during the International Polar Year 2007-09. Permafrost and Periglacial Processes, 21(2): 182-197 Vikeb, F.B., Huse, A., Slotte, A, Stenevik, E.K., and Lien, V.S., 2010: Effect of hatching date, vertical distribution and interannual variation in physical forcing on northward displacement and temperature conditions of Norwegian spring-spawning herring larvae. ICES Journal of Marine Science, 67, 1948-1956. Vikhamar-Schuler, D., I. Hanssen-Bauer, T.V., Schuler, S.D. Mathiesen, and M. Lehning. 2013.: Use of a multilayer snow model to assess grazing conditions for reindeer. Annals of Glaciology. 54(62) pp 214-226. Vincent, W., J. Hobbie, J. Laybourn-Parry, 2008: Introduction to the Limnology of High-latitude Lake and River Ecosystems. In: Polar lakes and Rivers: Limnology of Arctic and Antarctic Aquatic Ecosystems. [Laybourn- Parry, J. and J. Laybourn-Parry(eds.)]. Oxford University Press, pp. 1-18. Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Vincent, W.F., L.G. Whyte, C. Lovejoy, C.W. Greer, I. Laurion, C.A. Suttle, J. Corbeil, and D.R. Mueller, 2009: Arctic microbial ecosystems and impacts of extreme warming during the International Polar Year. Polar Science, 3(3), 171-180. Virginia, R.S. and K.S. Yalowitz, 2012: A New Paradigm for Arctic Health: Challenges and Responses to Rapid Climate, Environmental, and Social Change. In: Workshop Report for May 23-25, Applied Circumpolar Policy [Virginia, R.S. and Yalowitz, K.S. (eds.)]. Dartmouth College, Hanover, N.H. Voggesser, G., K. Lynn, J. Daigle, F. Lake, D. Ranco. 2013 Cultural impacts to tribes from climate change influences on forests. Special Issue Climatic Change. Vol. 12. Issue 3. 615-626. Von Biela, V., R., C. E. Zimmerman, and L. L. Moulton, 2011. Long-term increases in young-of-the-year growth of Arctic cisco Coregonus autumnalis and environmental influences. Journal Fish Biology 78(1): 39-56. Vors, L.S. and M.S. Boyce, 2009: Global declines of caribou and reindeer. Global Change Biology, 15(11), 2626- 2633. Vuojala-Magga, T., M. Turunen, T. Ryyppö, and M. Tennberg, 2011: Resonance Strategies of Sámi Reindeer Herders in Northernmost Finland during Climatically Extreme Years. Arctic, 64(2). Vyverman, W., E. Verleyen, A. Wilmotte, D.A. Hodgson, A. Willems, K. Peeters, B. Van de Vijver, A. De Wever, F. Leliaert, and K. Sabbe, 2010: Evidence for widespread endemism among Antarctic micro-organisms. Polar Science, 4(2), 103-113. 10.1016/j.polar.2010.03.006. Walker, D.A., M.O. Leibman, H.E. Epstein, B.C. Forbes, U.S. Bhatt, M.K. Raynolds, J.C. Comiso, A.A. Gubarkov, A.V. Khomutov, G.J. Jia, E. Kaarlejärvi, J.O. Kaplan, T. Kumpula, P. Kuss, G. Matyshak, N.G. Moskalenko, P. Orekhov, V.E. Romanovsky, N.G. Ukraientseva, and Q. Yu, 2009: Spatial and temporal patterns of greenness on the Yamal Peninsula, Russia: Interactions of ecological and social factors affecting the Arctic normalized difference vegetation index. Environmental Research Letters, 4(4), 16. Walker, D.A., P. Kuss, H.E. Epstein, A.N. Kade, C.M. Vonlanthen, M.K. Raynolds, and F.J. Daniëls, 2011: Vegetation of zonal patterned-ground ecosystems along the North America Arctic bioclimate gradient. Applied Vegetation Science, 14(4), 440-463. Walsh, J.J., D.A. Dieterle, F.R.Chen, J.M. Lenes, W. Maslowski, J.J. Cassano, T.E. Whitledge, D. Stockwell, M. Flint, R.N. Sukhanova, J. Christensen. 2011. Trophic cascades and future harmful algal blooms within ice-free Arctic seas north of Bering Strait. A simulation analysis. Progress in Oceanography. 91 (2011) 312-343. http://www.sciencedirect.com/science/article/pii/S0079661111000164 Walter, K.M., J.P. Chanton, Chapin,F.S.,,III, E.A.G. Schuur, and S.A. Zimov, 2008: Methane production and bubble emissions from arctic lakes: Isotopic implications for source pathways and ages. 113, G00A08. Walter, K.M., L.C. Smith, and F. Stuart Chapin, 2007a: Methane bubbling from northern lakes: present and future contributions to the global methane budget. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 365(1856), 1657-1676. Walter, K.M., M.E. Edwards, G. Grosse, S.A. Zimov and F.S. Chapin III, 2007b. Thermokarst lakes as a source of atmospheric CH4 during the last deglaciation. Science, 318:633-636. Waluda, C.M., M.A. Collins, A.D. Black, I.J. Staniland, and P.N. Trathan, 2010: Linking predator and prey behaviour: contrasts between Antarctic fur seals and macaroni penguins at South Georgia. Marine Biology, 157(1), 99-112. Walvoord, M.A. and R.G. Striegl, 2007: Increased groundwater to stream discharge from permafrost thawing in the Yukon River basin: Potential impacts on lateral export of carbon and nitrogen. Geophysical Research Letters, 34(12), L12402. Wang and Overland 2011. I do not have access to the literature databasae and cannot search this ref. But it was in the Geophysical Research Letters Wasley, J., S.A. Robinson, M. Popp, and C.E. Lovelock, 2006: Some like it wet biological characteristics underpinning tolerance of extreme water events in Antarctic bryophytes. Functional Plant Biology, 33, 443- 455. Wassman, P., C. M. Duarte, S. Agusti, and M. K. Sejr, 2011: Footprints of climate change in the Arctic marine ecosystem. Global Change Biology, 17,1235-1249. Wassmann, P., 2011: Arctic marine ecosystems in an era of rapid climate change. Progress in Oceanography, 90, 1- 17. Watson, S., L.S. Peck, P.A. Tyler, P.C. Southgate, K.S. Tan, R.W. Day, and S.A. Morley, 2012: Marine invertebrate skeleton size varies with latitude, temperature and carbonate saturation: implications for global change and ocean acidification. Global Change Biology, 18(10), 3026-3038. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Weatherhead, E., S. Gearheard, and R.G. Barry, 2010: Changes in weather persistence: Insight from Inuit knowledge. Global Environmental Change, 20(3), 523-528. Weimerskirch, H., M. Louzao, S. de Grissac, and K. Delord, 2012: Changes in wind pattern alter albatross distribution and life-history traits. Science, 335, 211-214. Weimerskirch, H., P. Inchausti, C. Guinet, and C. Barbraud, 2003: Trends in bird and seal populations as indicators of a system shift in the Southern Ocean. Antarctic Science, 15(2), 249-256. Wenzel, G.W., 2009: Canadian Inuit subsistence and ecological instability? if the climate changes, must the Inuit? Polar Research, 28(1), 89-99. West, J. and G. Hovelsrud, 2010: Cross-scale Adaptation Challenges in the Coastal Fisheries: Findings from Lebesby, Northern Norway. Arctic, 63(3). Whyte, K.P. 2013: Justice forward: tribes, climate adaptation and responsibility. Special Issue Climatic Change. Vol. 12, Issue 3. 517-530. Wiedenmann, J., K. Cresswell, and M. Mangel, 2008: Temperature-dependent growth of Antarctic krill: predictions for a changing climate from a cohort model. Mar Ecol Prog Ser, 358, 191-202. Wiedenmann, J., K.A. Cresswell, and M. Mangel, 2009: Connecting recruitment of Antarctic krill and sea ice. Limonology and Oceanography, 54(3), 799-811. Wiig, O., J. Aars, and E.W. Born, 2008: Effects of climate change on polar bears. Science Progress, 91(Pt 2), 151- 173. Wildcat, D. 2009. Red Alert!: Saving the Planet with Indigenous Knowledge. Golden, CO: Fulcrum Publishing. 143 pp. Wilderbuer, T., W. Stockhausen, N. Bond, 2013: Updated analysis of flatfish recruitment response to climate variability and ocean conditions in the Eastern Bering Sea. Deep-Sea Research II, 94, 157-164. Williams, T. and P. Hardison. 2013. Culture, law, risk and governance: contexts of traditional knowledge inm climate change adaptation. Special Issue Climatic Change. Vol. 12. Issue 3. 531d-544. Williams, T.M., S.R. Noren and M. Glenn. 2011. Extreme physiological adaptations as predictors of climate-change sensitivity in the narwhal, Monodon monoceros. Mar. Mamm. Sci., 27, 334-349. Wilson, W.J.,and O.A.Ormseth, 2009: A new management plan for Arctic waters of the United States. Fisheries, 34(11), 555-558. Wolf, A., T.V. Callaghan, and K. Larson, 2008: Future changes in vegetation and ecosystem function of the Barents Region. Climatic Change, 87(1-2), 51-73. Wramneby, A., B. Smith, and P. Samuelsson, 2010: Hot spots of vegetation-climate feedbacks under future greenhouse forcing in Europe. Journal of Geophysical Research D: Atmospheres, 115(21), 12. Xu, L., R.B. Myneni, F.S. Chapin III, T.V. Callaghan, J.E. Pinzon, C.J. Tucker, Z. Zhu, J. Bi, P. Ciais, H. Tommervik, E.S. Euskirchen, B.C. Forbes, S.L. Piao, B.T. Anderson, S. Ganguly, R.R. Nemani, S.J. Goetz, P.S.A. Beck, A.G. Bunn, C. Cao, and J.C. Stroeve, 2013: Temperature and vegetation seasonality diminishment over northern lands. Nature Climate Change, advance online publication. Yamamoto-Kawai, M., F.A. McLaughlin, E.C. Carmack, S. Nishino, and K. Shimada, 2009: Aragonite undersaturation in the Arctic Ocean: Effects of ocean acidification and sea ice melt. Science, 326(5956), 1098- 1100. Yarie, J., 2008: Effects of moisture limitation on tree growth in upland and floodplain forest ecosystems in interior Alaska. Forest Ecology and Management, 256(5), 1055-1063. Ye, B., D. Yang, Z. Zhang, and D.L. Kane, 2009: Variation of hydrological regime with permafrost coverage over Lena Basin in Siberia. 114, D07102. Zarnetske, J.P., M.N. Gooseff, W.B. Bowden, M.J. Greenwald, T.R. Brosten, J.H. Bradford, and J.P. McNamara, 2008: Influence of morphology and permafrost dynamics on hyporheic exchange in arctic headwater streams under warming climate conditions. Geophysical Research Letters, 35(2), L02501. Zeng H., Jia G., Forbes B., 2013. Shifts in Arctic phenology in response to climate and anthropogenic factors as detected from multiple satellite time series. Environmental Research Letters 8, Zerbini, A., P. Clapham, and P. Wade, 2010: Assessing plausible rates of population growth in humpback whales from life-history data. Marine Biology, 157(6), 1225-1236, DOI: 10.1007/s00227-010-1403-y. Zhang, J., Y.H. Spitz, M. Steele, C. Ashjian, R. Campbell, L. Berline, and P. Matrai, 2010: Modeling the impact of declining sea ice on the Arctic marine planktonic ecosystem. 115, C10015. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Zhang, K., J.S. Kimball, E.H. Hogg, M. Zhao, W.C. Oechel, J.J. Cassano, and S.W. Running, 2008: Satellite-based model detection of recent climate-driven changes in northern high-latitude vegetation productivity. Journal of Geophysical Research G: Biogeosciences, 113(3), 13. Zhang, X., J. He, J. Zhang, I. Polyakov, I., R. Gerdes, J. Inoue and P. Wu, 2013. Enhanced poleward moisture transport and amplified high-latitude wetting trend. Nature Climate Change, 3: 47-51 Zhukova, N.G., V.N. Nesterova, I.P. Prokopchuk, and G.B. Rudneva, 2009: Winter distribution of euphausiids (Euphausiacea) in the Barents Sea (2000 2005). Deep Sea Research Part II: Topical Studies in Oceanography, 56(21 22), 1959-1967. Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 28-1: Annually averaged changes in inland and maritime transportation accessibility by mid-century (2045 2059) versus baseline (2000 2014). Change in winter road- Change in maritime-accessible accessible land area (km2) ocean area (km2) (Type A (2,000 kg GVWRvehicle) vessel) current EEZ Canada -13% 19% Finland -41% 0% Greenland -11% 28% Iceland -82% <1% Norway -51% 2% Russia -13% 16% Sweden -46% 0% USA (Alaska) -29% 5% High seas n/a 406% Total -14% 23% Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 28-1: Location maps of the north and south polar regions. Credit: P. Fretwell, British Antarctic Survey. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 28-2: Temporal change in onset of flowering (plants), median date of emergence (arthropods) and clutch initiation dates (birds) in high-Arctic Greenland. Red dots are statistically significant, blue dots are not. Source: Hye et al., 2007. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 28-3: Significant changes (p< 0.01) in photosynthetically active period NDVI between 1982 and 2012. Source: Xu et al., 2013. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 28 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 28-4: Projected duration of the navigation period (days) over the Northwest Passage and Northern Sea Route. Source: Mokhow and Khon, 2008. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 29. Small Islands Coordinating Lead Authors Leonard Nurse (Barbados), Roger McLean (Australia) Lead Authors John Agard (Trinidad and Tobago), Lino Pascal Briguglio (Malta), Virginie Duvat (France), Netatua Pelesikoti (Samoa), Emma Tompkins (UK), Arthur Webb (Fiji) Contributing Authors John Campbell (New Zealand), Dave Chadee (Trinidad and Tobago), Shobha Maharaj (Trinidad and Tobago), Veronique Morin (Canada), Geert Jan van Oldenborgh (Netherlands), Rolph Payet (Seychelles), Daniel Scott (Canada) Review Editors Thomas Spencer (UK), Kazuya Yasuhara (Japan) Volunteer Chapter Scientist Veronique Morin (Canada) Contents Executive Summary 29.1. Introduction 29.2. Major Conclusions from Previous Assessments 29.3. Observed Impacts of Climate Change, including Detection and Attribution 29.3.1. Observed Impacts on Island Coasts and Marine Biophysical Systems 29.3.1.1. Sea-Level Rise, Inundation, and Shoreline Change on Small Islands 29.3.1.2. Coastal Ecosystem Change on Small Islands: Coral Reefs and Coastal Wetlands 29.3.2. Observed Impacts on Terrestrial Systems: Island Biodiversity and Water Resources 29.3.3. Observed Impacts on Human Systems in Small Islands 29.3.3.1. Observed Impacts on Island Settlements and Tourism 29.3.3.2. Observed Impacts on Human Health 29.3.3.3. Observed Impacts of Climate Change on Relocation and Migration 29.3.3.4. Observed Impacts on Island Economies 29.3.4. Detection and Attribution of Observed Impacts of Climate Change on Small Islands 29.4. Projected Integrated Climate Change Impacts 29.4.1. Non-Formal Scenario-Based Projected Impacts 29.4.2. Projected Impacts for Islands based on Scenario Projections 29.4.3. RCP Projections and Implications for Small Islands 29.5. Inter- and Intra-Regional Trans-Boundary Impacts on Small Islands 29.5.1. Large Ocean Waves from Distant Sources 29.5.2. Trans-Continental Dust Clouds and their Impact 29.5.3. Movement and Impact of Introduced and Invasive Species across Boundaries 29.5.4. Spread of Aquatic Pathogens within Island Regions 29.5.5. Trans-Boundary Movements and Human Health 29.6. Adaptation and Management of Risks Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 29.6.1. Addressing Current Vulnerabilities on Small Islands 29.6.2. Practical Experiences of Adaptation on Small Islands 29.6.2.1. Building Adaptive Capacity with Traditional Knowledge, Technologies, and Skills on Small Islands 29.6.2.2. Addressing Risks on Small Islands 29.6.2.3. Working Collectively to Address Climate Impacts on Small Islands 29.6.2.4. Addressing Long-Term Climate Impacts and Migration on Small Islands 29.6.3. Barriers and Limits to Adaptation in Small Island Settings 29.6.4. Mainstreaming and Integrating Climate Change into Development Plans and Policies 29.7. Adaptation and Mitigation Interactions 29.7.1. Assumptions/Uncertainties Associated with Adaptation and Mitigation Responses 29.7.2. Potential Synergies and Conflicts 29.8. Facilitating Adaptation and Avoiding Maladaptation 29.9. Research and Data Gaps References Frequently Asked Questions 29.1: Why is it difficult to detect and attribute changes on small islands to climate change? 29.2: Why is the cost of adaptation to climate change so high in small islands? 29.3: Is it appropriate to transfer adaptation and mitigation strategies between and within small island countries and regions? Executive Summary Current and future climate-related drivers of risk for small islands during the 21st century include sea-level rise, tropical and extra-tropical cyclones, increasing air and sea surface temperatures, and changing rainfall patterns (high confidence, robust evidence, high agreement) [WGI 14, Table 29-1]. Current impacts associated with these changes confirm findings reported on small islands from the AR4 and previous IPCC assessments. The future risks associated with these drivers include loss of adaptive capacity [29.6.2.1, 29.6.2.3] and ecosystem services critical to lives and livelihoods in small islands [29.3.1, 29.3.2, 29.3.3]. Sea-level rise poses one of the most widely recognized climate change threats to low-lying coastal areas on islands and atolls (high confidence, robust evidence and high agreement) [29.3.1]. It is virtually certain that global mean sea-level rise rates are accelerating [WGI 13.2.2.1]. Projected increases to the year 2100 (RCP 4.5: 0.35m to 0.70m, WGI 13.5.1, Table 29-1) superimposed on extreme sea-level events (e.g. swell waves, storm surges, ENSO) present severe sea-flood and erosion risks for low-lying coastal areas and atoll islands (high confidence). Likewise, there is high confidence that wave over-wash of sea water will degrade fresh ground water resources [29.3.2] and that sea surface temperature rise will result in increased coral bleaching and reef degradation [29.3.1.2]. Given the dependence of island communities on coral reef ecosystems for a range of services including coastal protection, subsistence fisheries and tourism, there is high confidence that coral reef ecosystem degradation will negatively impact island communities and livelihoods. Given the inherent physical characteristics of small islands, the AR5 reconfirms the high level of vulnerability of small islands to multiple stressors, both climate and non-climate (high confidence, robust evidence, high agreement). However, the distinction between observed and projected impacts of climate change is often not clear in the literature on small islands (high agreement) [29.3]. There is evidence that this challenge can be partly overcome through improvements in baseline monitoring of island systems and downscaling of climate-model projections, which would heighten confidence in assessing recent and projected impacts [WGI 9.6, 29.3, 29.4, 29.9]. Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Small islands do not have uniform climate change risk profiles (high confidence). Rather their high diversity in both physical and human attributes and their response to climate-related drivers means that climate change impacts, vulnerability and adaptation will be variable from one island region to another and between countries in the same region [Fig 29.1; Table 29.3]. In the past, this diversity in potential response has not always been adequately integrated in adaptation planning. There is increasing recognition of the risks to small islands from climate-related processes originating well beyond the borders of an individual nation or island. Such trans-boundary processes already have a negative impact on small islands (high confidence, robust evidence, medium agreement). These include: airborne dust from the Sahara and Asia, distant-source ocean swells from mid- high latitudes, invasive plant and animal species and the spread of aquatic pathogens. For island communities the risks associated with existing and future invasive species and human health challenges are projected to increase in a changing climate [29.5.4]. Adaptation to climate change generates larger benefit to small islands when delivered in conjunction with other development activities, such as disaster risk reduction and community based approaches to development (medium confidence) [29.6.4]. Addressing the critical social, economic and environmental issues of the day, raising awareness and communicating future risks to local communities [29.6.3] will likely increase human and environmental resilience to the longer-term impacts of climate change [29.6.1, 29.6.2.3, Figure 29-5]. Adaptation and mitigation on small islands are not always trade-offs, but can be regarded as complementary components in the response to climate change (medium confidence). Examples of adaptation-mitigation inter- linkages in small islands include energy supply and use, tourism infrastructure and activities, and functions and services associated with coastal wetlands. The alignment of these sectors for potential emission reductions together with adaptation, offer co-benefits and opportunities in some small islands [29.7.2, 29.8]. Lessons learned from adaptation and mitigation experiences in one island may offer some guidance to other small island states, though there is low confidence in the success of wholesale transfer of adaptation and mitigation options when the local lenses through which they are viewed differ from one island state to the next, given the diverse cultural, socio- economic, ecological and political values [29.6.2, 29.8]. The ability of small islands to undertake adaptation and mitigation programs, and their effectiveness, can be substantially strengthened through appropriate assistance from the international community (medium confidence). However, caution is needed to ensure such assistance is not driving the climate change agenda in small islands, as there is a risk that critical challenges confronting island governments and communities may not be addressed . Opportunities for effective adaptation can be found by, for example, empowering communities and optimizing the benefits of local practices that have proven to be efficacious through time, and working synergistically to progress development agendas [29.8, 29.6.2.3, 29.6.3]. 29.1. Introduction It has long been recognized that greenhouse gas emissions from small islands are negligible in relation to global emissions, but that the threats of climate change and sea-level rise to small islands are very real. Indeed, it has been suggested that the very existence of some atoll nations is threatened by rising sea levels associated with global warming. Whilst such scenarios are not applicable to all small island nations, there is no doubt that on the whole the impacts of climate change on small islands will have serious negative effects especially on socio-economic conditions and bio-physical resources although impacts may be reduced through effective adaptation measures. The small islands considered in this chapter are principally sovereign states and territories located within the tropics of the southern and western Pacific Ocean, central and western Indian Ocean, the Caribbean Sea, and the eastern Atlantic off the coast of west Africa, as well as in the more temperate Mediterranean Sea. Although these small islands nations are by no means homogenous politically, socially, or culturally, or in terms of physical size and character or economic development, there has been a tendency to generalise about the potential impacts on small islands and their adaptive capacity. In this chapter we attempt to strike a balance between Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 identifying the differences between small islands as well as recognising that small islands tend to share a number of common characteristics that have distinguished them as a particular group in international affairs. Also in this chapter we reiterate some of the frequently voiced and key concerns relating to climate change impacts, vulnerability and adaptation whilst emphasising a number of additional themes that have emerged in the literature on small islands since the IPCC Fourth Assessment. These include the relationship between climate change policy, activities and development issues; externally generated trans-boundary impacts; and the implications of risk in relation to adaptation and the adaptive capacity of small island nations. 29.2. Major Conclusions from Previous Assessments Small islands were not given a separate chapter in the IPCC First Assessment (FAR) in1990 though they were discussed in the chapter on World Oceans and Coastal Zones (Tsyban et al., 1990). Two points were highlighted. First, that a 30-50 cm sea-level rise projected by 2050 would threaten low islands, and that a 1 m rise by 2100 would render some island countries uninhabitable (Tegart et al., 1990). Second, the costs of protection works to combat sea-level rise would be extremely high for small island nations. Indeed, as a percentage of GDP the Maldives, Kiribati, Tuvalu, Tokelau, Anguilla, Turks and Caicos, Marshall Islands and Seychelles were ranked among the ten nations with the highest protection costs in relation to GDP (Tsyban et al., 1990). Over twenty years later these two points continue to be emphasized. For instance, although small islands represent only a fraction of total global damage projected to occur due to a sea-level rise of 1.0 m by 2100 (SRES A1 scenario) the actual damage costs for the small island states is enormous in relation to the size of their economies with several small island nations being included in the group of ten countries with the highest relative impact projected for 2100 (Anthoff et al., 2010). The Second Assessment (SAR) in 1995 confirmed the vulnerable state of small islands, now included in a specific chapter titled Coastal Zones and Small Islands (Bijlsma et al., 1996). However, importantly the SAR recognized that both vulnerability and impacts would be highly variable between small islands and that impacts were likely to be greatest where local environments are already under stress as a result of human activities (Bijlsma et al., 1996). The report also summarized results from the application of a common methodology for vulnerability and adaptation analysis that gave new insights into the socio-economic implications of sea-level rise for small islands including: negative impacts on virtually all sectors including tourism, freshwater resources, fisheries and agriculture, human settlements, financial services and human health; protection is likely to be very costly; and, adaptation would involve a series of tradeoffs. It also noted that major constraints to adaptation on small islands included: lack of technology and human resource capacity, serious financial limitations, lack of cultural and social acceptability and uncertain political and legal frameworks. Integrated coastal and island management was seen as a way of overcoming some of these constraints. The Third Assessment (TAR) in 2001 included a specific chapter on Small Island States . In confirming previously identified concerns of small island states two factors were highlighted, the first relating to sustainability noting that with limited resources and low adaptive capacity, these islands face the considerable challenge of meeting the social and economic needs of their populations in a manner that is sustainable (Nurse et al., 2001). And the second, that there were other issues faced by small island states concluding that for most small islands the reality of climate change is just one of many serious challenges with which they are confronted (Nurse et al., 2001). Both of these themes are raised again and assessed in the light of recent findings in the present chapter. Until the Fourth Assessment (AR4) in 2007, sea-level rise had dominated vulnerability and impact studies of small island states. Whilst a broader range of climate change drivers and geographical spread of islands was included in the Small Islands chapter, Mimura et al. (2007) prefaced their assessment by noting that the number of independent scientific studies on climate change and small islands since the TAR had been quite limited and in their view the volume of literature in refereed international journals relating to small islands and climate change since publication of the TAR is rather less than that between the SAR in 1995 and TAR in 2001 (Mimura et al., 2007). Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Since AR4 the literature on small islands and climate change has increased substantially. A number of features distinguish the literature we review here from that included in earlier assessments. First, the literature appears more sophisticated and does not shirk from dealing with the complexity of small island vulnerability, impacts and adaptation or the differences between islands and island states. Second, and related to the first, the literature is less one-dimensional, and deals with climate change in a multidimensional manner as just one of several stressors on small island nations. Third, the literature also critiques some aspects of climate change policy, notably in relation to critical present-day development and security needs of small islands [29.3.3.1] as well as the possibility that some proposed adaptation measures may prove to be maladaptive [29.8]. Fourth, many initiatives have been identified in recent times that will reduce vulnerability and enhance resilience of small islands to on-going global change including improving risk knowledge and island resource management while also strengthening socio-economic systems and livelihoods (Hay, 2013). 29.3. Observed Impacts of Climate Change, including Detection and Attribution The distinction between observed impacts of climate change and projected impacts is often unclear in the small islands literature and discussions. Publications frequently deal with both aspects of impacts interchangeably, and use observed impacts from, for instance an extreme event, as an analogy to what may happen in the future due to climate change (e.g. Lo-Yat et al., 2011). The key climate and ocean drivers of change that impact small islands include variations in air and ocean temperatures, ocean chemistry, rainfall, wind strength and direction, sea-levels and wave climate and particularly the extremes such as tropical cyclones, drought and distant storm swell events. All have varying impacts, dependent on the magnitude, frequency, temporal and spatial extent of the event, as well as on the bio-physical nature of the island (Figure 29.1) and its social, economic and political setting. [INSERT FIGURE 29-1 HERE Figure 29-1: Representative tropical island typologies. From top-left: a young, active volcanic island (with altitudinal zonation) and limited living perimeter reefs (purple zone at outer reef edge), through to an atoll (centre bottom) and raised limestone island (bottom right) dominated by ancient reef deposits (brown + white fleck). Atolls have limited, low-lying land areas but well developed reef/lagoon systems. Islands composed of continental rocks are not included in this figure, but see Table 29-3.] 29.3.1. Observed Impacts on Island Coasts and Marine Biophysical Systems 29.3.1.1. Sea-Level Rise, Inundation, and Shoreline Change on Small Islands Sea-level rise poses one of the most widely recognized climate change threats to low-lying coastal areas (Church and White 2011; Cazenave and Llovel, 2010; Nicholls and Cazenave, 2010). This is particularly important in small islands where the majority of human communities and infrastructure is located in coastal zones with limited on- island relocation opportunities especially on atoll islands (Woodroffe, 2008) (Figure 29-1). Over much of the 20th Century, global mean sea level rose at a rate between 1.3 to 1.7 mm yr 1 and since 1993, at a rate between 2.8 to 3.6 mm yr 1 (WGI, Table 13.1) and acceleration is detected in longer records since 1870 (Merrifield et al., 2009; Church and White, 2011; WGI, 13.2.2.1). Rates of sea-level rise are however not uniform across the globe and large regional differences have been detected including in the Indian Ocean and tropical Pacific, where in some parts rates have been significantly higher than the global average (Meyssignac et al., 2012) (refer also to 5.3.2.2). In the tropical western Pacific where a large number of small island communities exist, rates up to four times the global average (approximately 12 mm yr -1) have been reported between 1993 and 2009. These are generally thought to describe short-term variations associated with natural cyclic climate phenomena such as ENSO (El Nino-Southern Oscillation) which has a strong modulating effect on sea level variability with lower/higher-than-average sea level during El Nino/La Nina events of the order of +/-20-30 cm (Becker et al., 2012; Cazenave and Remy, 2011). Large interannual variability in sea level has also been demonstrated from the Indian Ocean (e.g. Chagos Archipelago, Dunne et al., 2012) whilst Palanisamy et al. (2012) found that over the last 60 years the mean rate of sea-level rise in the Caribbean region was similar to the global average ~ 1.8mm yr-1. Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 There are few long-term sea-level records available for individual small island locations. Reported sea flooding and inundation is often associated with transient phenomena, such as storm waves and surges, deep ocean swell and predicted astronomical tidal cycles (Vassie et al., 2004; Zahibo et al., 2007; Komar and Allan, 2008; Haigh et al., 2011). For example, high spring tide floods at Fongafale Island, Funafuti Atoll, Tuvalu, have been well publicized and areas of the central portion of Fongafale are already below high spring tide level. However, rates of relative sea- level rise at Funafuti between 1950 2009 have been approximately three times higher than the global average (Becker et al., 2012) and saline flooding of internal low-lying areas occurs regularly, and is expected to become more frequent and extensive over time (Yamano et al., 2007). Documented cases of coastal inundation and erosion often cite additional circumstances such as vertical subsidence, engineering works, development activities or beach mining as the causal process. Four examples can be cited. First, on the Torres Islands, Vanuatu communities have been displaced due to increasing inundation of low-lying settlement areas due to a combination of tectonic subsidence and sea-level rise (Ballu et al., 2011). Second, on Anjouan Island, Comores in the Indian Ocean, Sinane et al. (2010) found beach aggregate mining was a major contributing factor influencing rapid beach erosion. Third, the intrinsic exposure of rapidly expanding settlements and agriculture in the low-lying flood prone Rewa Delta, Fiji is shown by Lata and Nunn (2012) to place populations in increasingly severe conditions of vulnerability to flooding and marine inundation. Fourth, Hoeke et al. (2013) describe a 2008 widespread inundation event that displaced some 63,000 people in Papua New Guinea and Solomon Islands alone. That event was primarily caused by remotely generated swell waves, and the severity of flooding was greatly increased by anomalously high regional sea levels linked with ENSO and on-going sea-level rise. Such examples serve to highlight that extreme events superimposed on a rising sea-level baseline are the main drivers that threaten the habitability of low-lying islands as sea levels continue to rise. Since the AR4 a number of empirical studies have documented historical changes in island shorelines. Historical shoreline position change over 20 60 years on 27 central Pacific atoll islands showed that total land area remained relatively stable in 43 per cent of islands, whilst another 43 per cent had increased in area, and the rest showed a net reduction in land area (Webb and Kench, 2010). Dynamic responses were also found in a four year study of 17 relatively pristine islands on two other central Pacific atolls in Kiribati by Rankey (2011) who concluded that sea- level rise was not likely to be the main influencing factor in these shoreline changes. Similarly in French Polynesia Yates et al. (2013) showed mixed shoreline change patterns over the last 40 50 years with examples of both erosion and accretion in the 47 atoll islands assessed. Sea-level rise did not appear to be the primary control on shoreline processes on these islands. On uninhabited Raine Island on the Great Barrier Reef, Dawson and Smithers (2010) also found that shoreline processes were dynamic but that island area and volume increased 6 per cent and 4 per cent respectively between 1967 and 2007. Overall, these studies of observed shoreline change on reef islands conclude that for rates of change experienced over recent decades normal seasonal erosion and accretion processes appear to predominate over any long-term morphological trend or signal at this time. Ford s (2013) investigation of Wotje Atoll, Marshall Islands also found shoreline variability between 1945 and 2010 but that overall accretion had been more prevalent than erosion up until 2004. From 2004 to the present 17 out of 18 islands became net erosive, potentially corresponding to the high sea levels in the region over the last 10 years. On the high tropical islands of Kauai and Maui, Hawaii, Romine and Fletcher (2013) found shoreline change was highly variable over the last century but that recently chronic erosion predominated with over 70% of beaches now being erosive. Finally, it is important to note the majority of these studies warn that: (1) past changes cannot be simply extrapolated to determine future shoreline responses; and (2) rising sea level will incrementally increase the rate and extent of erosion in the future. In many locations changing patterns of human settlement and direct impacts on shoreline processes present immediate erosion challenges in populated islands and coastal zones (Yamano et al., 2007; Storey and Hunter, 2010; Novelo-Casanova and Suarez, 2010) and mask attribution to sea-level rise. A study of Majuro atoll (Marshall Islands) found that erosion was widespread but attribution to sea-level rise was obscured by pervasive anthropogenic impacts to the coastal system (Ford, 2012) (see also 5.4.4). Similarly a study of three islands in the Rosario Archipelago (Colombia) reported shoreline retreat over a 50-55 year period and found Grande, Rosario, and Tesoro Islands had lost 6.7, 8.2 and 48.7 per cent of their land area respectively. Erosion was largely attributed to poor management on densely settled Grande Island, whilst sea-level rise and persistent northeast winds enhanced erosion on uninhabited Rosario and Tesoro (Restrepo et al., 2012). Likewise, Cambers (2009) reported average beach Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 erosion rates of 0.5 m yr-1 in eight Caribbean islands from 1985-2000. Whilst the study could not quantify the extent of attribution it noted that greater erosion rates were positively correlated with the number of hurricane events. Alternately, Etienne and Terry (2012) found a Category 4 tropical cyclone that passed within 30 km of Taveuni Island (Fiji) nourished shorelines with fresh coralline sediments despite localized storm damage. Whilst these studies contribute to improved understanding of island shoreline processes and change since AR4, the warning of increased vulnerability of small island shores and low-lying areas to inundation and erosion in response to sea-level rise and other potential climate change stressors is not diminished. 29.3.1.2. Coastal Ecosystem Change on Small Islands: Coral Reefs and Coastal Wetlands Coral reefs are an important resource in small tropical islands and wellbeing of many island communities is linked to their on-going function and productivity. Reefs play a significant role in supplying sediment to island shores and in dissipating wave energy thus reducing the potential foreshore erosion. They also provide habitat for a host of marine species upon which many island communities are dependent for subsistence foods as well as underpinning beach and reef-based tourism and economic activity (Bell et al., 2011; Perch-Nielsen, 2010). The documented sensitivity of coral reef ecosystems to climate change is summarised elsewhere (see Chap 5, Box CC-CR). Increased coral bleaching and reduced reef calcification rates due to thermal stress and increasing CO2 concentration are expected to affect the functioning and viability of living reef systems (Hoegh-Guldberg et al., 2007; Eakin et al., 2009). Some studies already implicate thermal stress in reduced coral calcification rates (Tanzil et al., 2009) and regional declines in calcification of corals that form reef framework (De'ath et al., 2009; Cantin et al., 2010). Unprecedented bleaching events have been recorded in the remote Phoenix Islands (Kiribati) with nearly 100 per cent coral mortality in the lagoon and 62 per cent mortality on the outer leeward slopes of the otherwise pristine reefs of Kanton Atoll during 2002 / 2003 (Alling et al., 2007). Similar patterns of mortality were observed in four other atolls in the Phoenix group and temperature-induced coral bleaching was also recorded in isolated Palmyra Atoll during the 2009 ENSO event (Williams et al., 2010). In 2005 extensive bleaching was recorded at 22 sites around Rodrigues Island in the western Indian Ocean with up to 75 per cent of the dominant species affected in some areas (Hardman et al., 2007). Studies of the severe 1998 El Nino bleaching event in the tropical Indian Ocean showed reefs in the Maldives, Seychelles and Chagos Islands were among the most impacted (Cinner et al., 2012; Tkachenko, 2012). In 2005 a reef survey around Barbados following a Caribbean regional bleaching event revealed the most severe bleaching ever recorded with approximately 70 per cent of corals impacted (Oxenford et al., 2008). Globally, the incidence and implications of temperature-related coral bleaching in small islands is well documented and combined with the effects of increasing ocean acidification these stressors could threaten the function and persistence of island coral reef ecosystems (see Chap 5, Box CC-OA). Island coral reefs have limited defences against thermal stress and acidification. However studies such as Cinner et al. (2012) and Tkachenko (2012) highlight that whilst recovery from bleaching is variable, some reefs show greater resilience than others. There is also some evidence to show that coral reef resilience is enhanced in the absence of other environmental stresses such as declining water quality. In Belize chronologies of growth rates in massive corals (Montastraea faveolata) over the past 75 150 years suggest that the bleaching event in 1998 was unprecedented and its severity appeared to stem from reduced thermal tolerance related to human coastal development (Carilli et al., 2010). Likewise a study over a 40 year period (1960s 2008) in the Grand Recif of Tulear, Madagascar concluded that severe degradation of the reef was mostly ascribed to direct anthropogenic disturbance, despite an average 1 0C increase in temperature over this period (Harris et al., 2010). Coral recovery following the 2004 bleaching event in the central Pacific atolls of Tarawa and Abaiang (Kiribati) was also noted to be improved in the absence of direct human impacts (Donner et al., 2010) and isolation of bleached reefs was shown by Gilmour et al. (2013) to be less inhibiting to reef recovery than direct human disturbance. The loss of coral reef habitat has detrimental implications for coastal fisheries (Pratchett et al., 2009) in small islands where reef-based subsistence and tourism activities are often critical to the wellbeing and economies of islands (Bell et al., 2011). In Kimbe Bay, Papua New Guinea 65 per cent of coastal fish are dependent on living reefs at some stage in their life cycle and that following degradation of the reef fish abundance declined (Jones et al., 2004). Even where coral reef recovery has followed bleaching, reef associated species composition may not recover Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 to its original state (Pratchett et al., 2009; Donner et al., 2010). Sea Surface Temperature (SST) anomaly events can be associated with a lag in the larval supply of coral reef fishes, as reported by Lo-Yat et al. (2011) between 1996 and 2000 at Rangiroa Atoll, French Polynesia. Higher temperatures have also been implicated in negatively affecting the spawning of adult reef species (Munday et al., 2009; Donelson et al., 2010). Like coral reefs, mangroves and sea grass environments provide a range of ecosystem goods and services (Polidoro et al., 2010; Waycott et al., 2009) and both habitats play a significant role in the wellbeing of small island communities. Mangroves in particular serve a host of commercial and subsistence uses as well as providing natural coastal protection from erosion and storm events (Ellison, 2009; Krauss et al., 2010; Waycott et al., 2011). Sea-level rise is reported as the most significant climate change threat to the survival of mangroves (Waycott et al., 2011). Loss of the seaward edge of mangroves at Hungry Bay, Bermuda has been reported by Ellison (1993) who attributes this process to sea-level rise and the inability of mangroves to tolerate increased water depth at the seaward margin. Elsewhere in the Caribbean and tropical Pacific observations vary in regards to the potential for sedimentation rates in mangroves forests to keep pace with sea-level rise (McKee et al., 2007; Krauss et al., 2003). In Kosrae and Pohnpei Islands (Federated States of Micronesia), Krauss et al. (2010) found significant variability in mangrove average soil elevation changes due to deposition from an accretion deficit of 4.95 mm y-1 to an accretion surplus of 3.28 mm y-1 relative to the estimated rate of sea-level rise. Such surpluses are generally reported from high islands where additional sediments can be delivered from terrestrial runoff. However, Rankey (2011) described natural seaward migration (up to 40m) of some mangrove areas between 1969 and 2009 in atolls in Kiribati suggesting sediment accretion can also occur in sediment rich reefal areas and in the absence of terrigenous inputs. The response of seagrass to climate change is also complex, regionally variable and manifest in quite different ways. A study of seven species of sea grasses from tropical Green Island, Australia highlighted the variability in response to heat and light stress (Campbell et al., 2006). Light reduction may be a limiting factor to sea grass growth due to increased water depth and sedimentation (Ralph et al., 2007). Ogston and Field (2010) observed that a 20 cm rise in sea-level may double suspended sediment loads and turbidity in shallow waters on fringing reefs of Molokai, Hawaiian Islands, with negative implications to photosynthetic species such as seagrass. Otherwise, temperature stress is most commonly reported as the main expected climate change impact on seagrass (e.g. Campbell et al., 2006; Waycott et al., 2011). Literature on seagrass diebacks in small islands is scarce but research in the Balearic Islands (Western Mediterranean) has shown that over a six-year study, seagrass shoot mortality and recruitment rates were negatively influenced by higher temperature (Marbá and Duarte, 2010). (See also Chapter 5.4.2.3 for further discussion of impacts on mangrove and sea grass communities). 29.3.2. Observed Impacts on Terrestrial Systems: Island Biodiversity and Water Resources Climate change impacts on terrestrial biodiversity on islands, frequently interacting with several other drivers (Blackburn et al., 2004; Didham et al., 2005), fall into three general categories namely: (a) ecosystem and species horizontal shifts and range decline; (b) altitudinal species range shifts and decline mainly due to temperature increase on high islands; and (c) exotic and pest species range increase and invasions mainly due to temperature increase in high latitude islands. Due to the limited area and isolated nature of most islands, these effects are generally magnified compared to continental areas and may cause species loss especially in tropical islands with high numbers of endemic species. For example, in two low-lying islands in the Bahamas, Greaver and Sternberg (2010) found that during periods of reduced rainfall the shallow freshwater lens subsides and contracts landward and ocean water infiltrates further inland negatively impacting on coastal strand vegetation. Sea-level rise has also been observed to threaten the long-term persistence of freshwater-dependent ecosystems within low-lying islands in the Florida Keys (Goodman et al., 2012). On Sugarloaf Key, Ross et al. (2009) found pine forest area declined from 88 to 30 h from 1935 to 1991 due to increasing salinisation and rising ground water, with vegetation transitioning to more saline tolerant species such as mangroves. Whilst there are many studies that report observations associated with temperature increases in mid- and high- latitude islands, such as the Falkland Islands and Marion Islands in the south Atlantic and south Indian ocean respectively (Bokhorst et al., 2007, 2008; Le Roux et al., 2005) and Svalbard in the Arctic (Webb et al., 1998) there Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 are few equivalent studies in tropical small islands. A recent study of the tropical Mauritius kestrel indicate changing rainfall conditions in Mauritius over the last 50 years have resulted in this species having reduced reproductive success due to a mismatch between the timing of breeding and peak food abundance (Senapathi et al., 2011). Increasing global temperatures may also lead to altitudinal species range shifts and contractions within high islands with an upward creep of the tree line and associated fauna (Benning et al., 2002; Krushelnycky et al., 2013). A study in the Hawaiian Islands which assessed data from 21 stations over 85 years showed a rapid rise in surface temperatures over the last 30 years with stronger warming in mountain areas (CCSP, 2008). Comparative vegetation distribution and composition studies in sub-Antarctic Marion Island, found an altitudinal shift of 3.4 m yr -1 for plant species (Parolo and Rossi, 2008). Comparable effects also occur in the tropics such as in Hawaii Volcano National Park where comparison of sample plots over a 40 year period from 1966-67 to 2008 show fire-adapted grasses expanded upward along a warming tropical elevation gradient (Angelo and Daehler, 2013). Reduction in the numbers and sizes of endemic populations caused by such habitat constriction and changes in species composition in mountain systems may result in the demise and possibly extinction of endemic species (Chen et al., 2009; Pauli et al., 2007; Sekercioglu et al., 2008; Krushelnycky et al., 2013). Altitudinal temperature change has also been reported to influence the distribution for disease vectors such as mosquitoes potentially threatening biota unaccustomed to such vectors (Freed et al., 2005; Atkinson and LaPointe, 2009). Freshwater supply in small island environments has always presented challenges and has been an issue raised in all previous IPCC reports. On high volcanic and granitic islands small and steep river catchments respond rapidly to rainfall events and watersheds generally have restricted storage capacity. On porous limestone and low atoll islands surface runoff is minimal and water rapidly passes through the substrate into the groundwater lens. Rainwater harvesting is also an important contribution to freshwater access and alternatives like desalination have had mixed success in small island settings due to operational costs (White and Falkland, 2010). Rapidly growing demand, land use change, urbanisation and tourism are already placing significant strain on the limited freshwater reserves in small island environments (Cashman et al., 2010; Emmanuel and Spence, 2009; White and Falkland, 2010). In the Caribbean, where there is considerable variation in the types of freshwater supplies utilised, concern over the status of freshwater availability has been expressed for at least the past 30 years (Cashman et al., 2010). There have also been economic and management failures in the water sector not only in the Caribbean (Mycoo, 2007) but also in small islands in the Indian (Payet and Agricole, 2006) and Pacific oceans (Moglia et al., 2008a, 2008b; White et al., 2007). These issues also occur on a background of decreasing rainfall and increasing temperature. Rainfall records averaged over the Caribbean region for 100 years (1900-2000) show a consistent 0.18 mm yr -1 reduction in rainfall, a trend that is projected to continue (Jury and Winter, 2010). In contrast, analysis of rainfall data over the past 100 years from the Seychelles has shown substantial variability related to ENSO. Nevertheless an increase in average rainfall from 1959 to 1997 and an increase in temperature of ~ 0.25 o C per decade have occurred (Payet and Agricole, 2006). Long-term reduction in streamflow (median reduction of 22 23%) has been detected in the Hawaiian Islands over the period 1913 2008, resulting in reduced freshwater availability for both human use and ecological processes (Bassiouni and Oki 2013). Detection of long-term statistical change in precipitation is an important prerequisite towards a better understanding the impacts of climate change in small island hydrology and water resources. There is a paucity of empirical evidence linking saline (sea-water) intrusion into fresh groundwater reserves due simply to incremental sea-level rise at this time (e.g. Rozell and Wong, 2010). However this dynamic must be the subject of improved research given the importance of groundwater aquifers in small island environments. White and Falkland s (2010) review of existing small island studies indicates that a sea-level increase of up to 1 m would have negligible salinity impacts on atoll island groundwater lenses so long as there is adequate vertical accommodation space, island shores remain intact, rainfall patterns do not change and direct human impacts are managed. However, wave overtopping and washover can be expected to become more frequent with sea-level rise and this has been shown to impact freshwater lenses dramatically. On Pukapuka Atoll, Cook Islands storm surge over-wash occurred in 2005. This caused the freshwater lenses to become immediately brackish and took 11 months to recover to conductivity levels appropriate for human use (Terry and Falkland, 2010). The ability of the freshwater lens to float Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 upwards within the substrate of an island in step with incremental sea-level rise also means that in low-lying and central areas of many atoll islands the lens may pond at the surface. This phenomenon already occurs in central areas of Fongafale Island, Tuvalu, and during extreme high king tides large areas of the inner part of the island become inundated with brackish waters (Yamano et al., 2007; Locke, 2009). 29.3.3. Observed Impacts on Human Systems in Small Islands 29.3.3.1. Observed Impacts on Island Settlements and Tourism Whilst traditional settlements on high islands in the Pacific were often located inland, the move to coastal locations was encouraged by colonial and religious authorities and more recently through the development of tourism (Barnett and Campbell, 2010). Now the majority of settlement, infrastructure and development are located on lowlands along the coastal fringe of small islands. In the case of atoll islands, all development and settlement is essentially coastal. It follows that populations, infrastructure, agricultural areas and fresh groundwater supplies are all vulnerable to extreme tides, wave and surge events and sea-level rise (Walsh et al., 2012). Population drift from outer islands or from inland, together with rapid population growth in main centres and lack of accommodation space drives growing populations into ever more vulnerable locations (Connell, 2012). Additionally, without adequate resources and planning, engineering solutions such as shoreline reclamation also place communities and infrastructure in positions of increased risk (Duvat, 2013; Yamano et al., 2007). Many of the environmental issues raised by the media relating to Tuvalu, the Marshall Islands and Maldives are primarily relevant to the major population centre and its surrounds, which are Funafuti, Majuro and Male respectively. As an example Storey and Hunter (2010) indicate the Kiribati problem does not refer to the whole of Kiribati but rather to the southern part of Tarawa atoll where pre-existing issues of severe overcrowding, proliferation of informal housing and unplanned settlement, inadequate water supply, poor sanitation and solid waste disposal, pollution and conflict over land ownership are of concern. They argue that these problems require immediate resolution if the vulnerability of the South Tarawa community to the real and alarming threat of climate change is to be managed effectively (Storey and Hunter, 2010). On Majuro atoll, rapid urban development and the abandonment of traditional settlement patterns has resulted in movement from less vulnerable to more vulnerable locations on the island (Spennemann, 1996). Likewise, geophysical studies of Fongafale Island, the capital of Tuvalu, show that engineering works during World War II, and rapid development and population growth since independence, has led to the settlement of inappropriate shoreline and swampland areas, leaving communities in heightened conditions of vulnerability (e.g. Yamano et al., 2007). Ascribing direct climate change impacts in such disturbed environments is problematic due to the existing multiple lines of stress on the island s biophysical and social systems. However, it is clear that such pre-existing conditions of vulnerability add to the threat of climate change in such locations. Increased risk can also result from lack of awareness, particularly in communities in rural areas and outer islands ( periphery ) of archipelagic countries such as Cook Islands, Fiji, Kiribati and Vanuatu, whose climate change knowledge often contrasts sharply with that of communities in the major centres ( core ). In the core, communities tend to be better informed and have higher levels of awareness about the complex issues associated with climate change than in the periphery (Nunn et. al., 2013). The issue of coastal squeeze remains a concern for many small islands as there is a constant struggle to manage the requirements for physical development against the need to maintain ecological balance (Fish et al., 2008; Gero et al., 2011; Mycoo, 2011). Martinique in the Caribbean exemplifies the point, where physical infrastructure prevents the beach and wetlands from retreating landward as a spontaneous adaptation response to increased rates of coastal erosion (Schleupner, 2008). Moreover, intensive coastal development in the limited coastal zone combined with population growth and tourism has placed great stress on the coast of some islands and has resulted in dense aggregations of infrastructure and people in potentially vulnerable locations. Tourism is an important weather and climate-sensitive sector on many small islands and has been assessed on several occasions, including in previous IPCC assessments. There is currently no evidence that observed climatic Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 changes in small island destinations or source markets have permanently altered patterns of demand for tourism to small islands, and the complex mix of factors that actually determines destination choices under a changing climate still need to be fully evaluated (Scott et al., 2012a). However, there are cases reported that clearly show severe weather-related events in a destination country (e.g. heavy, persistent rainfall in Martinique: Hubner and Gössling, 2012; Hurricanes in Anguilla: Forster et al., 2012) can significantly influence visitors perception of the desirability of the location as a vacation choice. Climate can also impact directly on environmental resources that are major tourism attractions in small islands. Widespread resource degradation challenges such as beach erosion and coral bleaching have been found to negatively impact the perception of destination attractiveness in various locations, for example in Martinique (Schleupner, 2008), Barbados and Bonaire (Uyarra et al., 2005). Similarly dive tourists are well aware of coral bleaching, particularly the experienced diver segment (Gössling et al., 2012a; Klint et al., 2012). Therefore more acute impacts are felt by tourism operators and resorts that cater to these markets. Houston (2002) and Buzinde et al. (2010) also indicate that beach erosion may similarly affect accommodation prices in some destinations. Consequently, some countries have begun to invest in a variety of resource restoration initiatives including artificial beach nourishment, coral and mangrove restoration and the establishment of marine parks and protected areas (McClanahan et al., 2008; Mycoo and Chadwick, 2012). There is no analysis of how widespread such investments are or their capability to cope effectively with future climate change. The tourism industry and investors are also beginning to consider the climate risk of tourism operations (Scott et al., 2012b) including those associated with the availability of freshwater. Freshwater is limited on many small islands, and changes in its availability or quality during drought events linked to climate change have adverse impacts on tourism operations (UNWTO 2012). Tourism is a seasonally significant water user in many island destinations and in times of drought, concerns over limited supply for residents and other economic activities become heightened (Gössling et al., 2012b). The increasing use of desalination plants is one adaptation to reduce the risk of water scarcity in tourism operations. 29.3.3.2. Observed Impacts on Human Health Globally, the effects of climate change on human health will be both direct and indirect, and are expected to exacerbate existing health risks, especially in the most vulnerable communities where the burden of disease is already high (refer to Chapter 11.3, 11.5 and 11.6.1, this volume). Many small island states currently suffer from climate-sensitive health problems, including morbidity and mortality from extreme weather events, certain vector- and food- and water-borne diseases (Lozano, 2006; Barnett and Campbell, 2010; Cashman et al.,2010; Pulwarty et al., 2010; McMichael and Lindgren, 2011). Extreme weather and climate events such as tropical cyclones, storm surges, flooding, and drought can have both short- and long-term effects on human health, including drowning, injuries, increased disease transmission, and health problems associated with deterioration of water quality and quantity. Most small island nations are in tropical areas with weather conducive to the transmission of diseases such as malaria, dengue, filariasis and schistosomiasis. The linkages between human health, climate variability and seasonal weather have been demonstrated in several recent studies. The Caribbean has been identified as a highly endemic zone for leptospirosis with Trinidad and Tobago, Barbados and Jamaica representing the highest annual incidence (12, 10 and 7.8 cases per 100,000 population) in the world with only the Seychelles being higher (43.2 per 100,000 population) (Pappas et al., 2008). Studies conducted in Guadeloupe demonstrated a link between El Nino occurrence and leptospirosis incidence with rates increasing to 13 per 100,000 population in El Nino years as opposed to 4.5 cases per 100,000 inhabitants in La Nina and neutral years (Herrmann-Storck et al., 2008). In addition, epidemiological studies conducted in Trinidad reviewed the incidence of leptospirosis during the period 1996-2007 and showed seasonal patterns in the occurrence of confirmed leptospirosis cases, with significantly (P<0.001) more cases occurring in the wet season, May to November (193 cases), than during the dry season, December to May (66 cases) (Mohan et al., 2009). Recently changes in the epidemiology of leptospirosis have been detected especially in tropical islands with the main factors being climatic and anthropogenic ones (Pappas et al., 2008). These factors may be enhanced with increases in ambient temperature and changes in precipitation, vegetation and water availability as a consequence of climate change (Russell, 2009). Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In Pacific islands the incidence of diseases such as malaria and dengue fever has been increasing, especially endemic dengue in Samoa, Tonga and Kiribati (Russell, 2009). While studies conducted so far in the Pacific have only established a direct link between malaria, dengue and climate variability, these and other health risks including from cholera, are projected to increase as a consequence of climate change (Russell, 2009; refer also to Chapter 11.2.4 and 11.2.5 this volume, for detailed discussion on the link between climate change and projected increases in the outbreak of dengue and cholera). Dengue incidence is also a major health concern in other small island countries, including Trinidad and Tobago, Singapore, Cape Verde, Comoros and Mauritius (Chadee 2009; Koh et al., 2008; Van Kleef et al., 2010; Teles, 2011). In the specific cases of Trinidad and Tobago and Singapore the outbreaks have been significantly correlated with rainfall and temperature, respectively (Chadee et al., 2007; Koh et al., 2008). Previous IPCC assessments have consistently shown that human health on islands can be seriously compromised by lack of access to adequate, safe, freshwater and adequate nutrition (Nurse et al., 2001; Mimura et al., 2007). Lovell (2011) notes that in the Pacific many of the anticipated health effects of climate change are expected to be indirect, connected to the increased stress and declining well-being that comes with property damage, loss of economic livelihood and threatened communities. There is also a growing concern in island communities in the Caribbean and Pacific and Indian oceans, that fresh water scarcity, more intense droughts and storms could lead to a deterioration in standards of sanitation and hygiene (Cashman et al., 2010; McMichael and Lindgren, 2011). In such circumstances, increased exposure to a range of health risks including communicable (transmissible) diseases would be a distinct possibility. Ciguatera fish poisoning (CFP) occurs in tropical regions and is the most common non-bacterial food-borne illness associated with consumption of fish. Distribution and abundance of the organisms that produce these toxins, chiefly dinoflagellates of the genus Gambierdiscus, are reported to correlate positively with water temperature. Consequently, there is growing concern that increasing temperatures associated with climate change could increase the incidence of CFP in the island regions of the Caribbean (Morrison et al., 2008; Tester et al., 2010), Pacific (Rongo and van Woesik, 2011; Chan et al., 2011), the Mediterranean (Aligizaki and Nikolaidis, 2008; refer also to section 29.5.5), and the Canary Islands in the Atlantic (Pérez-Arellano et al., 2005). A recent Caribbean study sought to characterise the relationship between sea surface temperatures (SSTs) and CFP incidence and to determine the effects of temperature on the growth rate of organisms responsible for CFP. Results from this work show that in the Lesser Antilles high rates occur in areas that experience the warmest water temperatures and which show the least temperature variability (Tester et al., 2010). There are also high rates in the Pacific in Tokelau, Tuvalu, Kiribati, Cook Islands and Vanuatu (Chan et al., 2011). The influence of climatic factors on malaria vector density and parasite development is well established (Béguin et al., 2011; Chaves and Koenraadt, 2010). Previous studies have assessed the potential influence of climate change on malaria, using deterministic or statistical models (Hay et al., 2009; Martens et al., 1999; Parham and Michael, 2010; Pascual et al., 2006). While the present incidence of malaria on small islands is not reported to be high, favourable environmental and social circumstances for the spread of the disease are present in some island regions and are expected to be enhanced under projected changes in climate in Papua New Guinea, Guyana, Suriname and French Guyana (Michon et al., 2007; Rawlins et al., 2008; Figueroa, 2008). In the Caribbean, the occurrence of autochthonous malaria in non-endemic island countries in the last ten years suggests that all of the essential malaria transmission conditions now exist. Rawlins et al. (2008) call for enhanced surveillance, recognizing the possible impact of climate change on the spread of the anopheles mosquito vector and malaria transmission. 29.3.3.3. Observed Impacts of Climate Change on Relocation and Migration Evidence of human migration as a response to climate change is scarce for small islands. While there is general agreement that migration is usually driven by multiple factors (Black et al., 2011), several authors highlight the lack of empirical studies of the effect of climate-related factors, such as sea-level rise, on island migration (Liller and Van den Broeck, 2011; Mortreux and Barnett, 2009). Furthermore, there is no evidence of any government policy that allows for climate refugees from islands to be accepted into another country (Bedford and Bedford, 2010). This finding contrasts with the early desk-based estimates of migration under climate change such as the work of Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Myers (2002). These early studies have been criticised as they fail to acknowledge the reality of climate impacts on islands, the capacity of islands and islanders to adapt, or the actual drivers of migration (Barnett and O'Neill, 2012). Studies of island migration commonly reveal the complexity of a decision to migrate and rarely identify a single cause. For example, when looking at historical process of migration within the Mediterranean it appears that rising levels of income, coupled with a decreased dependence on subsistence agriculture has left the Mediterranean less vulnerable to all environmental stressors, resulting in a reduced need for mobility to cope with environmental or climatic change (de Haas, 2011). Studies from the Pacific have also shown that culture, lifestyle and a connection to place are more significant drivers of migration than climate (Barnett and Webber, 2010). For example, a Pacific Access Category of migration has been agreed between New Zealand and Tuvalu that permits 75 Tuvaluans to migrate to New Zealand every year (Kravchenko, 2008). Instead of enabling climate driven migration, this agreement is designed to facilitate economic and social migration as part of the Pacific island lifestyle (Shen and Gemenne, 2011). To date there is no unequivocal evidence that reveals migration from islands is being driven by anthropogenic climate change. There is however some evidence that environmental change has played a role in Pacific Island migration in the past (Nunn, 2007). In the Pacific environmental change has been shown to affect land use and land rights, which in turn have become drivers of migration (Bedford and Bedford, 2010). In a survey of 86 case studies of community relocations in Pacific islands, Campbell et al. (2005) found that environmental variability and natural hazards accounted for 37 communities relocating. In the Pacific, where land rights are a source of conflict, climate change could increase levels of stress associated with land rights and impact on migration (Campbell, 2010; Weir and Virani, 2011). While there is not yet a climate fingerprint on migration and resettlement patterns in all small islands, it is clear that there is the potential for human movement as a response to climate change. To better understand the impact of climate change on migration there is an urgent need for robust methods to identify and measure the effects of the drivers of migration on migration and resettlement. 29.3.3.4. Observed Impacts on Island Economies The economic and environmental vulnerabilities of small islands states are well documented (Briguglio et al., 2009, Bishop, 2012). Such vulnerabilities, which render the states at risk of being harmed by economic and environmental conditions, stem from intrinsic features of these vulnerable states, and are not usually governance induced. However, governance does remain one of the challenges for island countries in the Pacific in the pursuit of sustainable development through economic growth (Prasad, 2008). Economic vulnerability is often the result of a high degree of exposure to economic conditions often outside the control of small island states, exacerbated by dependence on a narrow range of exports and a high degree of dependence on strategic imports, such as food and fuel (Briguglio et al., 2009). This leads to economic volatility, a condition that is harmful for the economy of the islands (Guillaumont, 2010). There are other economic downsides associated with small size and insularity. Small size leads to high overhead cost per capita, particularly in infrastructural outlays. This is of major relevance to climate change adaptation that often requires upgrades and redesign of island infrastructure. Insularity leads to high cost of transport per unit, associated with purchases of raw materials and industrial supplies in small quantities, and sales of local produced products to distant markets. These disadvantages are associated with the inability of small islands to reap the benefits of economies of scale resulting in a high cost of doing business in small islands (Winters and Martins, 2004). High costs are also associated with the small size of island states when impacted by extreme events such as hurricanes and droughts. On small islands such events often disrupt most of the territory, especially on single-island states, and have a very large negative impact on the state s GDP, in comparison with larger and more populous states where individual events generally only affect a small proportion of the country and have a small impact on its GDP (Anthoff et al., 2010). Moreover, the dependence of many small islands on a limited number of economic sectors such as tourism, fisheries and agricultural crops, all of which are climate-sensitive, means that on the one Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 hand climate change adaptation is integral to social stability and economic vitality but that government adaptation efforts are constrained because of the high cost on the other. 29.3.4. Detection and Attribution of Observed Impacts of Climate Change on Small Islands While exceptional vulnerability of many small islands to future climate change is widely accepted, the foregoing analysis indicates that the scientific literature on observed impacts is quite limited. Detection of past and recent climate change impacts is challenging due to the presence of other anthropogenic drivers, especially in the constrained environments of small islands. Attribution is further challenged by the strong influence of natural climate variability compared to gradual incremental change of climate drivers. Notwithstanding these limitations a summary of the relationship between detection and attribution to climate change of several of the phenomena described in the above sections has been prepared. Figure 29-2 reflects the degree of confidence in the linkage between observed changes in several components of the coastal, terrestrial and human systems of small islands and the drivers of climate change. [INSERT FIGURE 29-2 HERE Figure 29-2: A comparison of the degree of confidence in the detection of observed impacts of climate change on tropical small islands with the degree of confidence in attribution to climate change drivers at this time. For example, the blue symbol No. 2 (Coastal Systems), indicates there is very high confidence in both the detection of sea-level rise consistent with global means and its attribution to climate change drivers; whereas the red symbol No. 17 (Human Systems) indicates whilst detection of casualties and damage during extreme events is very high, there is presently low confidence in the attribution to climate change. It is important to note that low confidence in attribution frequently arises due to the limited research available on small island environments.] 29.4. Projected Integrated Climate Change Impacts Small islands face many challenges in using climate change projections for policy development and decision- making (Keener et al., 2012). Among these is the inaction inherent in the mismatch of the short-term time scale on which government decisions are generally taken compared with the long-term time scale required for decisions related to climate change. This is further magnified by the general absence of credible regional socio-economic scenarios relevant at the spatial scale at which most decisions are taken. Scenarios are an important tool to help decision makers disaggregate vulnerability to the direct physical impacts of the climate signal from the vulnerability associated with socio-economic conditions and governance. There is however a problem in generating formal climate scenarios at the scale of small islands since they are generally much smaller than the resolution of the global climate models. This is because the grid squares in the Global Circulation Models (GCMs) used in the SRES scenarios over the last decade, were between 200 and 600 km2 that provides inadequate resolution over the land areas of most small islands. This has recently improved with the new RCP scenario GCMs with grid boxes generally between 100 and 200 km2 in size. The scale problem has been usually addressed by the implementation of statistical downscaling models that relate GCM output to the historical climate of a local small island datapoint. The limitation of this approach is the need for observed data ideally for at least three decades for a number of representative points on the island, in order to establish the statistical relationships between GCM data and observations. In most small islands long-term quality- controlled climate data are generally sparse, so that in widely dispersed islands such as in the Pacific, observational records are usually supplemented with satellite observations combined with dynamical downscaling computer models (Australian Bureau of Meteorology [ABoM] and CSIRO, 2011a; Keener et al., 2012). However where adequate local data are available for several stations for at least 30 years, downscaling techniques have demonstrated that they can provide projections at fine scales ranging from about 10 25 km2 (e.g. Charlery and Nurse, 2010; ABoM and CSIRO, 2011a). Even so, most projected changes in climate for the Caribbean, Pacific, Indian Ocean and Mediterranean islands, generally apply to the region as a whole and this may be adequate to determine general trends in regions where islands are close together. Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 29.4.1. Non-Formal Scenario-Based Projected Impacts Scenarios are often constructed by using a qualitative or broad order of magnitude climate projections approach based on expected changes in some physical climate signal from literature review rather than projections based on direct location specific modelling. Usually this is proposed as a what if question which is then quantified using a numerical method. For example in the Pacific, digital elevation models of Fiji s islands have been used to identify high risk areas for flooding based on six scenarios for sea-level rise from 0.09 to 0.88 m in combination with six scenarios for storm surge with return intervals from 1 to 50 years (Gravelle and Mimura, 2008). Another example of qualitative modelling from the Pacific is a case study from Nauru which uses local data and knowledge of climate to assess the GCM projections. It suggests that Nauru should plan for continued ENSO variability in the future with dry years during La Nina and an overall increase in mean rainfall and extreme rainfall events. Climate adaptation concerns which arise include water security and potential changes in extreme wet events which affect infrastructure, and human health (Brown et al., 2013a). Climate change also poses risks for food security in the Pacific islands, including agriculture and fisheries (Barnett, 2011) Projections have also been used in the islands of the Republic of Baharain to estimate proness to inundation for sea-level rise of 0.5, 1.0 and 1.5 m (Al-Jeneid et al., 2008). Similarly, in the Caribbean the elevation equivalent of a projected sea-level rise of 1 m has been superimposed on topographic maps to estimate that 49-60% of tourist resort properties would be damaged, potentially transforming the competitive position and sustainability of coastal tourism destinations in the region (Scott et al., 2012c). This method has also been used to quantify the area loss for over 12 900 islands and over 3000 terrestrial vertebrates in the tropical Pacific region for three sea-level rise scenarios. The study estimated that for sea-level rise of 1 m, 37 island endemic species in this region risk complete inundation (Wetzel et al., 2013). 29.4.2. Projected Impacts for Islands based on Scenario Projections Another approach to scenario development is to use the region specific projections more directly. It is worth noting that the broad synthesis in the AR4 of medium emissions climate scenario projections for small island regions (Mimura et al., 2007) shows concordance with the new RCP scenarios (see Table 29-1 and new RCP projections in Figure 29-3). For example, the SRES A1B medium emissions scenario suggests about a 1.8 to 2.3 °C median annual increase in surface temperature in the Caribbean, Indian Ocean and Pacific Ocean small islands regions by 2100 compared to a 1980-1999 baseline, with an overall annual decrease in precipitation of about 12% in the Caribbean (Table 11.1 in AR4 WG1; AR5 WG 1, 14.7.4) and a 3-5 per cent increase in the Indian and Pacific Oceans small island regions. Comparative projections for the new RCP4.5 scenario suggests about a 1.2 to 2.3 °C increase in surface temperature by 2100 compared to a 1986-2005 baseline and a decrease in precipitation of about 5 or 6 per cent in the Caribbean and Mediterranean respectively signaling potential future problems for agriculture and water availability compared to a 1-9 per cent increase in the Indian Ocean and Pacific Ocean small islands regions (Table 29.1). However, there are important spatial and high-island topography differences. Thus for example, among the more dispersed Pacific islands where the equatorial regions are likely to get wetter and the sub-tropical high pressure belts drier (as reported by AR5 WG I) in regions directly affected by the South Pacific Convergent Zone (SPCZ) and western portion of the Inter-Tropical Convergent Zone (ITCZ), the rainfall outlook is uncertain (AR5 WG1, 14.7.13). Projections for the Mediterranean islands also differ from those for the tropical small islands. Throughout the Mediterranean region, the length, frequency, and/or intensity of warm spells or heat waves are very likely to increase to the year 2100 (AR5 WGI, 14.7.6). Sea-level rise projections in the small islands regions for RCP4.5 are similar to the global projections of 0.41 to 0.71m (WGI AR5 13.5.1) ranging from 0.5-0.6 m by 2100 compared to 1986-2005 in the Caribbean, Pacific and Indian Ocean to 0.4-0.5 m in the Mediterranean and North Indian Ocean (Table 29-1). [INSERT TABLE 29-1 HERE Table 29-1: Climate change projections for the intermediate low (500-700 ppm CO2-e) RCP4.5 scenario for six Small Islands regions. The table shows the 25th, 50th (median) and 75th percentiles for surface air temperature and precipitation based on averages from 42 CMIP5 global models (adapted from WGI AR5 Table 14.1). Mean net regional sea-level change is evaluated from 21 CMIP5 models and includes regional non-scenario components (adapted from WGI AR5 Figure 13-20).] Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In the main regions in which most tropical or sub-tropical small island states are located, there are few independent peer reviewed scientific publications providing downscaled climate data projections, and even less illustrating the experience gained from their use for policy making. A possible 2 °C temperature increase by the year 2100 has potentially far reaching consequences for sentinel ecosystems such as coral reefs that are important to tropical islands (see Chapter 6.2.2.4.4.). This is because Degree Heating Months (DHM) >2 °C-month are the determining threshold for severe coral bleaching (Donner, 2009). For example in a study of sea surface temperature (SST) across all coral reef regions using GCM ensemble projections forced with five different SRES future emissions scenarios, Donner (2009) concluded that even warming in the future from the current accumulation of greenhouse gases in the atmosphere could cause over half of the world s coral reefs to experience harmfully frequent thermal stress by 2080. Further, this timeline could be brought forward to as early as 2030 under the A1B medium emissions scenario. He further stated that thermal adaptation of 1.5 °C would only delay the thermal stress forecast by 50 80 years. Donner (2009) also estimated the year of likelihood of a severe mass coral bleaching event due more than once every 5 years, to be 2074 in the Caribbean, 2088 in the western Indian Ocean, 2082 in the central Indian Ocean, 2065 in Micronesia, 2051 in the central Pacific, 2094 in Polynesia and 2073 in the eastern Pacific small islands regions. Using the new RCP scenarios by comparison, van Hooidonk et al. (2013) found that the onset of annual bleaching conditions is associated with about 510 ppm CO2 equivalent. The conclusion based on outputs from a wide range of emissions scenarios and models is that preserving >10 per cent of coral reefs worldwide would require limiting warming to less than 1.5+/-1.3 °C compared to pre-industrial levels (Frieler et al., 2013). Small island economies can also be objectively shown to be at greater risk from sea-level rise in comparison to other geographic areas since most of their population and infrastructure are in the coastal zone. This is demonstrated in a study using the Climate Framework for Uncertainty, Negotiation and Distribution (FUND) model to assess the economic impact of substantial sea-level rise in a range of socio-economic scenarios downscaled to the national level, including the four SRES storylines (Anthoff et al., 2010). Although this study showed that in magnitude, a few regions will experience most of the absolute costs of sea-level rise by 2100, especially East Asia, North America, Europe and South Asia, these same results when expressed as percent of GDP showed that most of the top ten and four of the top five most impacted are small islands from the Pacific (Federated States of Micronesia, Palau, Marshall Islands, Nauru) and Caribbean (Bahamas). The point is made that the damage costs for these small island states are enormous in relation to the size of their economies (Nicholls and Tol, 2006) and that together with deltaic areas they will find it most difficult to locally raise the finances necessary to implement adequate coastal protection (Anthoff et al., 2010). In the Caribbean, downscaled climate projections have been generated for some islands using the Hadley Centre PRECIS regional model (Taylor et al., 2007; Stephenson et al., 2008). For the SRES A2 and B2 scenarios the PRECIS regional climate model projects an increase in temperature across the Caribbean of 1 4 °C compared to a 1960-1990 baseline, with increasing rainfall during the latter part of the wet season from November January, in the northern Caribbean (i.e. north of 22°N) and drier conditions in the southern Caribbean linked to changes in the Caribbean Low Level Jet (CLLJ) with a strong tendency to drying in the traditional wet season from June October (Whyte et al., 2008; Campbell et al., 2011; Taylor et al., 2013). Projected lengthening seasonal dry periods, and increasing frequency of drought are expected to increase demand for water throughout the region under the SRES A1B scenario (Cashman et al., 2010). Decrease in crop yield is also projected in Puerto Rico for the SRES B1 (low), A2 (mid-high) and A1F1 scenarios during September although increased crop yield is suggested during February (Harmsen et al., 2009). Using a tourism demand model linked to the SRES A1F1 A2 B1 and B2 scenarios, the projected climate change heating and drying impacts are also linked to potential aesthetic, physical and thermal effects that are estimated to cause a change in total regional tourist expenditure of about +321, +356, -118 and -146 million US dollars from the least to the most severe emissions scenario respectively (Moore, 2010). In the Indian Ocean, representative downscaled projections have been generated for Australia s two Indian Ocean territories, the Cocos (Keeling) Islands and Christmas Island using the CSIRO Mark 3.0 climate model with the SRES A2 high emissions scenario (Maunsell Australia Pty Ltd., 2009). Future climate change projections for the two islands for 2070 include an approximate 1.8 °C increase in air temperature by 2070, probable drier dry seasons and wet seasons, about a 40 cm rise in sea-level and a decrease in the number of intense tropical cyclones. Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In the western tropical Pacific, extensive climate projections have been made for several Pacific Island Countries based on downscaling from an ensemble of models (ABoM and CSIRO, 2011b). The temperature projections in this region dominated by oceans seem less than those seen globally, ranging from +1.5 to 2.0 °C for the B1 low emissions scenario to +2.5 to 3.0 °C for the A2 high emissions scenario by the year 2090 relative to a 20 year period centred on 1990. Notably, extreme rainfall events that currently occur once every 20 years on average are generally simulated to occur four times per 20-year period, on average, by 2055 and seven times per 20-year period, on average, by 2090 under the A2 (high emissions) scenario (ABoM and CSIRO, 2011b). The results are not very different from the tropical Pacific RCP4.5 projections with projected temperature increases of about +1.2 to 1.4 °C by 2100 and an increase in rainfall of about 4% (Table 29-1). A comprehensive assessment of the vulnerability of the fisheries and aquaculture sectors to climate change in 22 Pacific island countries and territories focused on two future time-frames (2035 and 2100) and two SRES emissions scenarios, B1 (low emissions) and A2 (high emissions) (Bell et al., 2013). Many anticipated changes in habitat and resource availability such as coral reef-based fisheries are negative. By contrast, projected changes in tuna fisheries and freshwater aquaculture/fisheries can be positive with implications for government revenue and island food security (Bell et al., 2013). Simulation studies on changes in stocks of skipjack and bigeye tuna in the tropical Pacific area summarized in Table 29-2 and also discussed in 7.4.2.1 and 30.6.2.1.1. Some of these projected changes may favour the large international fishing fleets that can shift operations over large distances compared to local, artisanal fishers (Polovina et al., 2011). [INSERT TABLE 29-2 HERE Table 29-2: Summary of projected percentage changes in tropical Pacific tuna catches by 2035 and 2100 relative to 1980-2000 and the estimated resulting percentage changes to government revenue (After Bell et al., 2011).] In the Mediterranean islands of Mallorca, Corsica, Sardinia, Crete and Lesvos, Gritti et al. (2006) simulated the terrestrial vegetation biogeography and distribution dynamics under the SRES A1F1 and B1 scenarios to the year 2050. The simulations indicate that the effects of climate change are expected to be negligible within most ecosystems except for mountainous areas. These areas are projected to be eventually occupied by exotic vegetation types from warmer, drier conditions. Cruz et al. (2009) report similar results for the terrestrial ecosystems of Madeira Island in the Atlantic. Downscaled SRES A2 and B2 scenarios for the periods 2040 2069 and 2070 2099 suggest that the higher altitude native humid forest called the Laurissilva, may expand upwards in altitude, which could lead to a severe reduction of the heath woodland which because it has little upward area to shift may reduce in range or disappear at high altitudes resulting in the loss of rare and endemic species within this ecosystem. 29.4.3. RCP Projections and Implications for Small Islands Utilizing updated historical greenhouse gas emissions data the scientific community has produced future projections for four plausible new global Representative Concentration Pathways (RCPs) in order to explore a range of global climate signals up to the year 2100 and beyond (e.g. Moss et al., 2010). Typical model ensemble representations of low, intermediate low, intermediate high and high RCP projections for annual temperature and precipitation in some small islands regions are presented in Figure 29-3. Highlighted in Figure 29-3 is the ensemble mean of each RCP. A more comprehensive compilation of quarterly global RCP projections can be found in the Annex I Atlas of Global and Regional Climate Projections in the WGI AR5 Report. [INSERT FIGURE 29-3 HERE Figure 29-3: Time series of RCP scenarios annual projected temperature and precipitation change relative to 1986- 2005 for six small islands regions (using regions defined in AR5 WG1, Annex 1: Atlas of Global and Regional Climate Projections).] Thin lines denote one ensemble member per model, thick lines the CMIP5 multi-model mean. On the right-hand side the 5th, 25th, 50th (median), 75th and 95th percentiles of the distribution of 20-yr mean changes are given for 2081 2100 in the four RCP scenarios. Note that the model ensemble averages in the figure are for grid points over wide areas and encompass many different climate change signals. To get projections for a specific location and time period use the maps in the Atlas or the online interactive version at but please note that in regions with small islands the models basically simulate the climate of the surrounding ocean and local conditions on land may differ.] Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 During negotiations towards a new multi-lateral climate change regime Small Island Developing States (SIDS) have advocated that any agreement should be based on Global Mean Surface Temperature (GMST) increase well below 1.5 °C above pre-industrial levels (Hare et al., 2011; Riedy and McGregor, 2011). Inspection of column 1 in Figure 29-3 suggests that for the Caribbean, Indian Ocean and Pacific SIDS in the tropics, the median projected regional increase is in the range 0.5-0.9 °C by 2100 compared to 1986-2005. This together with the temperature change that has already occurred since the industrial revolution suggests that a temperature well below 1.5 °C is unlikely to be achieved with the lowest RCP2.6 projection (Peters et al., 2013). By comparison temperature projections for the intermediate low RCP4.5 scenario, Table 29-1 and Figure 29-3 suggest possible 1.2-1.5 °C temperature increases in Caribbean, Indian Ocean and Pacific SIDS by 2100 compared to 1986-2005. Similarly, the projections for the Mediterranean would be about a 2.3 °C increase by 2100 compared to 1986-2005 that would represent a 2.7 °C increase compared to pre-industrial temperatures. Associated with this change, the Caribbean and Mediterranean regions may experience a noticeable decrease in mean rainfall while the Indian and Pacific Ocean SIDS may experience increased rainfall. These trends accelerate moderately for RCP 6.0 and steeply for RCP 8.5 (Table 29-1). 29.5. Inter- and Intra-Regional Trans-Boundary Impacts on Small Islands Available literature since AR4 has highlighted previously less well understood impacts on small islands that are generated by processes originating in another region or continent well beyond the borders of an individual archipelagic nation or small island. These are inter-regional trans-boundary impacts. Intra-regional trans-boundary impacts originate from a within-region source (e.g. within the Caribbean). Some trans-boundary processes may have positive effects on the receiving small island or nation, though most that are reported have negative impacts. Deciphering a climate change signal in inter- and intra-regional trans-boundary impacts on small islands is not easy and usually involves a chain of linkages tracing back from island-impact to a distant climate or climate-related bio- physical or human process. Some examples are given below. 29.5.1. Large Ocean Waves from Distant Sources Unusually large deep ocean swells, generated from sources in the mid- and high-latitudes by Extra-Tropical Cyclones (ETC) cause considerable damage on the coasts of small islands thousands of kilometres away in the tropics. Impacts include sea-flooding and inundation of settlements, infrastructure and tourism facilities as well as severe erosion of beaches (see also 5.4.3.4). Examples from small islands in the Pacific and Caribbean are common though perhaps the most significant instance, in terms of a harbinger of climate change and sea-level rise, occurred in the Maldives in April 1987 when long period swells originating from the Southern Ocean some 6000 km away caused major flooding, damage to property, destruction of sea defences and erosion of reclaimed land and islands (Harangozo, 1992). The Maldives and several other island groups in the Indian Ocean have been subject to similar ocean swell events more recently, most notably in May 2007 (Department of Meteorology, 2007). In the Caribbean, northerly swells affecting the coasts of islands have been recognized as a significant coastal hazard since the 1950s (Donn and McGuinness, 1959). They cause considerable seasonal damage to beaches, marine ecosystems and coastal infrastructure throughout the region (Cambers, 2009; Bush et al., 2009). These high-energy events manifest themselves as long period high-amplitude waves that occur during the northern hemisphere winter and often impact the normally sheltered, low-energy leeward coasts of the islands. Such swells have even reached the shores of Guyana on the South American mainland as illustrated by a swell event in October 2005 that caused widespread flooding and overtopping and destruction of sea defences (van Ledden et al., 2009). Distant origin swells differ from the normal wave climate conditions experienced in the Caribbean, particularly with respect to direction of wave approach, wave height and periodicity and in their morphological impact (Cooper et al., 2013). Swells of similar origin and characteristics also occur in the Pacific (Fletcher et al., 2008; Keener et al., 2012). These events frequently occur in the Hawaiian island where there is evidence of damage to coral growth by swell from the north Pacific, especially during years with a strong El Nino signal (Fletcher et al., 2008). Hoeke et al. (2013) describe inundation from mid-high latitude north and south Pacific waves respectively at Majuro (Marshall Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Islands) in November and December 1979 and along the Coral Coast (Fiji) in May 2011. They also describe in detail an inundation event in December 2008 that was widespread throughout the western and central Pacific and resulted in waves surging across low-lying islands causing severe damage to housing and infrastructure and key natural resources that affected about 100,000 people across the region. The proximate cause of this event was swell generated in mid-latitudes of the North Pacific Ocean, more than 4000 km from the furthest affected island (Hoeke et al., 2013). Whereas the origin of the long period ocean swells that impact small islands in the tropical regions come from the mid-and high-latitudes in the Pacific, Indian and Atlantic oceans, there are also instances of unusually large waves generated from tropical cyclones that spread into the mid- and high- latitudes. One example occurred during 1999 when tide gauges at Ascension and St. Helena Islands in the central south Atlantic recorded unusually large deep- ocean swell generated from distant Hurricane Irene (Vassie et al., 2004). The impacts of increasing incidence or severity of storms or cyclones is generally considered from the perspective of direct landfall of such systems, whereas all of these instances serve to show the potential importance of swells to communities on distant, low-lying coasts, particularly if the climatology of swells is modified under future climate change (Vassie et al., 2004). From the perspective of those islands that suffer damage from this coastal hazard on an annual basis, this is an area that warrants further investigation. Projected changes in global wind-wave climate to 2070-2100, compared to a base period 1979-2009, show considerable regional and seasonal differences with both decreases and increases in annual mean significant wave height. Of particular relevance in the present context is the projected increase in wave activity in the Southern Ocean which influences a large portion of the global ocean as swell waves propagate northwards into the Pacific, Indian and Atlantic oceans (Hemer et al., 2013). Deep ocean swell waves and elevated sea-levels resulting from ETCs are examples of inter-regional trans-boundary processes; locally generated Tropical Cyclones (TCs) provide examples of intra-regional trans-boundary processes. Whilst hurricane force winds, heavy rainfall and turbulent seas associated with TCs can cause massive damage to both land and coastal systems in tropical small islands, the impacts of sea waves and inundation associated with far distant ETCs are limited to the coastal margins. Nevertheless both storm types result in a range of impacts covering island morphology, natural and ecological systems, island economies, settlements and human well-being (see Figure 29-4). [INSERT FIGURE 29-4 HERE Figure 29-4: Tropical and extra-tropical cyclone impacts on the coasts of small islands. Four types of impacts are distinguished here, black arrows showing the connections between them, based on the existing literature. An example of the chain of impacts associated with two extra-tropical cyclones centred to the east of Japan is illustrated by the red arrows. Swell waves generated by these events in December 2008 reached islands in the southwest Pacific and caused extensive flooding (3) that impacted soil quality (8), freshwater resources (9), and damaged crops (10), buildings (15), and transport facilities (16) in the region (Example based on Hoeke et al., 2013). Examples of tropical cyclone impacts on small island coasts with reference 1. Society Islands, French Polynesia, February 2010 (Etienne, 2012); 2. Taveuni, Fiji, March 2010 (Etienne and Terry, 2012); 3. Cook Islands (de Scally, 2008); Society and Autral Islands, French Polynesia, February 2010 (Etienne, 2012); 4. Viti Levu, Fiji, March 1997 (Terry et al., 2002); 5. Society Islands, French Polynesia, February 2010 (Etienne, 2012); 6. Curacao, Bonaire, Netherlands Antilles, November 1999 (Scheffers and Scheffers, 2006); Hawaiian Islands (Fletcher et al., 2008); 7. Bay Islands, Honduras, October 1998 (Cahoon et al., 2003); 8. Marshall Islands, June 1905 (Spennemann, 1996); 9. Pukapuka atoll, Cook Islands, February 2005 (Terry and Falkland, 2010); 10. Vanuatu, February 2004 (Richmond and Sovacool, 2012); 11. 12. 13. Tuamotu Islands, French Polynesia, 1982-83 (Dupon, 1987); 14. Grenada, September 2004 (OECS, 2004); 15. Grenada, September 2004 (OECS, 2004); Tubuai, Austral Islands, French Polynesia, February 2010 (Etienne, 2012); 16. Vanuatu, February 2004 (Richmond and Sovacool, 2012); Guadeloupe Island, October 2008 (Dorville and Zahibo, 2010); 17. Bora Bora, Raiatea, Maupiti, Tahaa, Huahine, Society Islands, February 2010 (Etienne, 2012); 18. Vanuatu, February 2004 (Richmond and Sovacool, 2012); 19. Tuamotu, French Polynesia, 1982-83 (Dupon, 1987). Examples of extra-tropical cyclone impacts on small island coasts with reference 1. Maldives, April 1987 (Harangozo, 1992); 2. Maldives, January 1955 (Maniku, 1990); 3. Maldives, April 1987 (Harangozo, 1992); 9. Solomon Islands, December 2008 (Hoeke et al., 2013); 10. Chuck, Pohnpei, Kosrae, Federated States of Micronesia, December 2008 (Hoeke et al., 2013); 15. Majuro, Marshall Islands, November 1979 Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (Hoeke et al., 2013); 16. Coral Coast, Viti Levu, Fiji, May 2011 (Hoeke et al., 2013); 17. Majuro, Kwajalein, Arno, Marshall Islands, December 2008 (Hoeke et al., 2013); 18. Bismark Archipelago, Papua New Guinea, December 2008 (Hoeke et al., 2013).] 29.5.2. Trans-Continental Dust Clouds and their Impact The transport of airborne Saharan dust across the Atlantic and into the Caribbean has engaged the attention of researchers for some time. The resulting dust clouds are known to carry pollen, microbes, insects, bacteria, fungal spores and various chemicals and pesticides (Prospero et al., 2005; Middleton et al., 2008; Monteil, 2008; López- Villarrubia et al., 2010; Garrison et al., 2006). During major events, dust concentrations can exceed 100 ug m-3 (Prospero, 2006). Independent studies using different methodologies have all found a strong positive correlation between dust levels in the Caribbean and periods of drought in the Sahara, while concentrations show a marked decrease during periods of higher rainfall. Consequently, it is argued that higher dust emissions due to increasing aridity in the Sahel and other arid areas could enhance climate change effects over large areas, including the eastern Caribbean and the Mediterranean (Prospero and Lamb, 2003). Similar findings have been reported at Cape Verde where dust emission levels were found to be a factor of nine lower during the decade of the 1950s when rainfall was at or above normal, compared to the 1980s, a period of intense drought in the Sahel region (Nicoll et al., 2011). Dust from the Sahara has also reached the eastern Mediterranean (e.g. Santese et al., 2010) whilst dust from Asia has been transported across the Pacific and Atlantic oceans and around the world (Uno et al., 2009). There is also evidence that the trans-boundary movement of Saharan dust into the island regions of the Caribbean, Pacific and Mediterranean is associated with various human health problems (Griffin, 2007) including asthma admissions in the Caribbean (Monteil, 2008; Monteil and Antoine, 2009; Prospero et al., 2008), cardiovascular morbidity in Cyprus in the Mediterranean (Middleton et al., 2008) and is found to be a risk factor in respiratory and obstructive pulmonary disease in the Cape Verde islands (Martins et al., 2009). These findings underscore the need for further research into the link between climate change, airborne aerosols and human health in localities such as oceanic islands far distant from the continental source of the particulates. 29.5.3. Movement and Impact of Introduced and Invasive Species across Boundaries Invasive species are coloniser species that establish populations outside their normal distribution ranges. The spread of invasive alien species is regarded as a significant trans-boundary threat to the health of biodiversity and ecosystems, and has emerged as a major factor in species decline, extinction and loss of biodiversity goods and services worldwide. This is particularly true of islands, where both endemicity and vulnerability to introduced species tend to be high (Kenis et al., 2009; Reaser et al., 2007; Westphal et al., 2008; Rocha et al., 2009; Kueffer et al., 2010). The extent to which alien invasive species successfully establish themselves at new locations in a changing climate will be dependent on many variables, but non-climate factors such as ease of access to migration pathways, suitability of the destination, ability to compete and adapt to new environments, and susceptibility to invasion of host ecosystems are deemed to be critical. This is borne out for example by Le Roux et al. (2008) who studied the effect of the invasive weed Miconia calvescens in New Caledonia, Society Islands and Marquesas Islands; by Gillespie et al. (2008) in an analysis of the spread of Leucaena leucocephala, Miconia calvescens, Psidium sp. and Schinus terebinthifolius in the Hawaiian islands; and by Christenhusz and Toivonen (2008) who showed the potential for rapid spread and establishment of the oriental vessel fern, Angiopteris evecta, from the South Pacific throughout the tropics. Mutualism between an invasive ant and locally honeydew-producing insects has been strongly associated with damage to the native and functionally important tree species, Pisonia grandis on Cousine Island, Seychelles (Gaigher, et al., 2011). Whilst invasive alien species constitute a major threat to biodiversity in small islands, the removal of such species can result in recovery and return of species richness. This has been demonstrated in Mauritius by Baider and Florens (2011) where some forested areas were weeded of alien plants and after a decade the forest had recovered close to its initial condition, They concluded, given the severity of alien plant invasion in Mauritius, that their example can Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 be seen as a relevant model for a whole swath of other island nations and territories around the world particularly in the Pacific and Indian Oceans (Baider and Florens, 2011). The movement of aquatic and terrestrial invasive fauna within and across regions will almost certainly exacerbate the threat posed by climate change in island regions, and could impose significant environmental, economic and social costs. Recent research has shown that the invasion of the Caribbean Sea by the Indo-Pacific lionfish (Pterois volitans), a highly efficient and successful predator, is a major contributor to observed increases in algal dominance in coral and sponge communities in the Bahamas and elsewhere in the region. The consequential damage to these ecosystems has been attributed to a significant decline in herbivores due to predation by lionfish (Schofield, 2010; Albins and Hixon, 2008; Lesser and Slattery, 2011; Green et al., 2011). While there is no evidence that the lionfish invasion is climate-related, the concern is that when combined with pre-existing stress factors the natural resilience of Caribbean reef communities will decrease (Albins and Hixon, 2013; Green et al., 2012), making them more susceptible to climate change effects such as bleaching. Englund (2008) has documented the negative effects of invasive species on native aquatic insects on Hawai i and French Polynesia, and their potential role in the extirpation of native aquatic invertebrates in the Pacific. Similarly, there is evidence that on the island of Oahu introduced slugs appear to be skewing species abundance in favour of certain non-native and native plants , by altering the rank order of seedling survival rates , thereby undermining the ability of preferred species (e.g. the endangered C. Superba) to compete effectively (Joe and Daehler, 2008). 29.5.4. Spread of Aquatic Pathogens within Island Regions The mass mortality of the black sea urchin, Diadema antillarum, in the Caribbean basin during the early 1980s demonstrates the ease with which ecological threats in one part of a region can be disseminated to other jurisdictions thousands of kilometres away. The die-off was first observed in the waters off Panama around January 1983, and within 13 months the disease epidemic had spread rapidly through the Caribbean Sea affecting practically all island reefs, as far away as Tobago some 2000 km to the south and Bermuda, some 4000 km to the east. The diadema population in the wider Caribbean declined between 90-95 per cent as a consequence of this single episode (Lessios, 1988, 1995) As D. antillarum is one of the principal grazers that removes macroalgae from reefs and thus promotes juvenile coral recruitment, the collateral damage was severe, as the region s corals suffered from high morbidity and mortality for decades thereafter (Carpenter and Edmunds, 2006; Idjadi et al., 2010). There are other climate-sensitive diseases such as yellow, white and black band, white plague and white pox that travel across national boundaries and infect coral reefs directly. This is variously supported by examples from the Indo-Pacific and Caribbean relating to the role of bacterial infections in white syndrome and yellow band disease (Piskorska et al., 2007; Cervino et al., 2008), the impact of microbial pathogens as stressors on benthic communities in the Mediterranean associated with warming seawater (Danovaro et al., 2009, and an increasing evidence of white, yellow and black band disease associated with Caribbean and Atlantic reefs (Rosenberg et al., 2009; Brandt and McManus, 2009; Miller et al., 2009a; Weil and Croquer, 2009; Weil and Rogers, 2011). 29.5.5. Trans-Boundary Movements and Human Health For island communities the trans-boundary implications of existing and future human health challenges are projected to increase in a changing climate. For instance, the aggressive spread of the invasive giant African snail, Achatina fulica, throughout the Caribbean, Indo-Pacific islands and Hawai i is not only assessed to be a severe threat to native snails and other fauna (e.g. native gastropods), flora and crop agriculture, but is also identified as a vector for certain human diseases such as meningitis (Reaser et al., 2007; Meyer et al., 2008; Thiengo et al., 2010). Like other aquatic pathogens, ciguatoxins that cause ciguatera fish poisoning may be readily dispersed by currents across and within boundaries in tropical and sub-tropical waters. Ciguatoxins are known to be highly temperature- sensitive and may flourish when certain sea water temperature thresholds are reached, as has been noted in the South Pacific (Llewellyn, 2010), Cook Islands (Rongo and van Woesik, 2011), Kiribati (Chan et al., 2011), the Caribbean Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and Atlantic (Otero et al., 2010; Tester et al., 2010) and Mediterranean (Aligizaki and Nikolaidis, 2008) (see also 29.3.3.2). 29.6. Adaptation and Management of Risks Islands face risks from both climate-related hazards that have occurred for centuries, as well as new risks from climate change. There have been extensive studies of the risks associated with past climate-related hazards and adaptations to these, such as tropical cyclones, drought, and disease, and their attendant impacts on human health, tourism, fisheries and other areas (Bijlsma et al., 1996; Cronk 1997; Solomon and Forbes 1999; Pelling and Uitto 2001). There have also been many studies that have used a variety of vulnerability, risk and adaptation assessment methods particularly in the Pacific that have recently been summarized by Hay et al., (2013). But for most islands, there is very little published literature documenting the probability, frequency, severity or consequences of climate change risks such as sea-level rise, ocean acidification, and salinisation of freshwater resources or associated adaptation measures. Projections of future climate change risks are limited by: the lack of model skill in projecting the climatic variables that matter to small islands, notably: tropical cyclone frequency and intensity, wind speed and direction, precipitation, sea-level, ocean temperature and ocean acidification (Brown et al., 2013b); inadequate projections of regional sea levels (Willis and Church, 2012), and a lack of long term baseline monitoring of changes in climatic risk, or to ground-truth models (Voccia, 2012), such as risk of saline intrusion, risk of invasive species, risk of biodiversity loss, or risk of large ocean waves. In their absence, qualitative studies have documented perceptions of change in current risks (Fazey et al., 2011; Lata and Nunn, 2012), reviewed effective coping mechanisms for current stressors (Bunce et al. 2009; Campbell et al., 2011) and have considered future scenarios of change (Weir and Virani 2011). These studies highlight that change is occurring, but they do not quantify the probability, speed, scale or distribution of future climate risks. The lack of quantitative published assessments of climate risk for many small islands means that future adaptation decisions have to rely on analogues of responses to past and present weather extremes and climate variability, or assumed/hypothesised impacts of climate change based on island type (Table 29-3). Differences in island type and differences in exposure to climate forcing and hazards vary with island form that provides a framework for consideration of vulnerability and adaptation strategies. Critical is a place-based understanding of island landscapes and of processes operating on individual islands (Forbes et al., 2013). [INSERT TABLE 29-3 HERE Table 29-3: Summary of island types in the Pacific region and implications for hydro-meteorological hazards.] 29.6.1. Addressing Current Vulnerabilities on Small Islands Islands are heterogeneous in geomorphology, culture, ecosystems, populations and hence also in their vulnerability to climate change Vulnerabilities and adaptation needs are as diverse as the variety of islands between regions and even within nation states (e.g. in Solomon Islands, Rasmussen et al., 2011), often with little climate adaptation occurring in peripheral islands, for example in parts of the Pacific (Nunn et al. 2013). Quantitative comparison of vulnerability is difficult due to the paucity of vulnerability indicators. Generic indices of national level vulnerability continue to emerge (Cardona, 2007) but only a minority are focused on small islands (e.g. Blancard and Hoarau, 2013). The island-specific indicators that exist often suffer from lack of data (Hughes et al., 2012; Peduzzi et al., 2009), use indicators that are not relevant in all islands (Barnett and Campbell, 2010), or use data of limited quality for islands, such as sea-level rise (as used in Wheeler, 2011). As a result indicators of vulnerability for small islands often misrepresent actual vulnerability Recent moves towards participatory approaches that link scientific knowledge with local visions of vulnerability (see Park et al., 2012) offers an important way forward to understanding island vulnerability in the absence of certainty in model-based scenarios. Island vulnerability is often a function of four key stressors: socio-economic, physical, socio-ecological and climate- induced , whose reinforcing mechanisms are important in determining the magnitude of impacts. Socio-economic vulnerabilities are related to on-going challenges of managing urbanisation, pollution and sanitation, both in small island states and non-sovereign islands as highlighted by Storey and Hunter (2010) in Kiribati, López-Marrero and Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Yarnal (2010) in Puerto Rico, and in Mayotte (France) (Le Masson and Kelman, 2011). Geo-physical characteristics of islands (see Table 29-2; Figure 29-1) create inherent physical vulnerabilities. Thus, for example the Azores (Portugal) face seismic, landslide and tsunami risks (Coutinho et al., 2009). Socio-ecological stresses, such as habitat loss and degradation, invasive species (described in Sax and Gaines, 2008), overexploitation, pollution, human encroachment and disease can harm biodiversity (Caujape-Castells et al., 2010; Kingsford et al., 2009), and reduce the ability of socio-ecological systems to bounce back after shocks. To understand climate vulnerability on islands, it is necessary to assess all of these dimensions of vulnerability (Rasmussen et al., 2011). For example, with individual ecosystems, such as coral reef ecosystems those already under stress from non-climate factors are more at risk from climate change than those that are unstressed (Maina et al., 2011; Hughes et al., 2003). Evidence is starting to emerge that shows the same applies at the island scale. In Majuro atoll (Marshall Islands), 34-37 years of aerial photography shows that socio-ecological stress is exacerbating shoreline change associated with sea-level rise, especially on the lagoon-side of islands (Ford, 2012; see also 29.3.1.1). Islands faced with multiple stressors can therefore be assumed to be more at risk from climate impacts. Despite the limited ability of continental scale models to predict climate risks for specific islands, or the limited capacity of island vulnerability indicators, scenario based damage assessments can be undertaken. Storm surge risks have been effectively modeled for the Andaman and Nicobar Islands (Kumar et al., 2008). Rainfall induced landslide risk maps have been produced for both Jamaica (Miller et al., 2009b) and the Chuuk Islands (Federated States of Micronesia) (Harp et al., 2009). However the probability of change in frequency and severity of extreme rainfall events and storm surges remains poorly understood for most small islands. Other risks, such as the climate change driven health risks from the spread of infectious disease, loss of settlements and infrastructure, and decline of ecosystems that affect island economies, livelihoods and human well-being also remain under-researched. Nevertheless, it is possible to consider these risks along with the threat of rising sea level and suggest a range of contemporary and future adaptation issues and prospects for small islands (Table 29-4). [INSERT TABLE 29-4 HERE Table 29-4: Selected key risks and potential for adaptation for small islands from the present-day to the long-term.] 29.6.2. Practical Experiences of Adaptation on Small Islands There is disagreement about whether islands and islanders have successfully adapted to past weather variability and climate change. Nunn (2007) argues that past climate changes have had a crisis effect on prehistoric societies in much of the Pacific Basin. In contrast a variety of studies argue that past experiences of hydro-meteorological extreme events have enabled islands to become resilient to weather extremes (Barnett, 2001). Resilience appears to come from both a belief in their own capacity (Kuruppu and Liverman, 2011; Adger and Brown, 2009), and a familiarity with their environment and understanding of what is needed to adapt (Tompkins et al., 2009; Le Masson and Kelman, 2011). For example, compared to communities in the larger countries of Madagascar, Tanzania and Kenya, the Indian Ocean islands the Seychelles and Mauritius were found to have: comparatively high capacity to anticipate change and prepare strategies; self-awareness of human impact on environment; willingness to change occupation; livelihood diversity; social capital; material assets; access to technology and infrastructure, all of which produced high adaptive capacity (Cinner et al., 2012). Despite this resilience, islands are assumed to be generically vulnerable to long term future climate change (Parks and Roberts, 2006; Myers, 2002). There are many ways in which in-situ climate adaptation can be undertaken: reducing socio-economic vulnerabilities, building adaptive capacity, enhancing disaster risk reduction, or building longer term climate resilience (e.g. see McGray et al., 2007; Eakin et al., 2009). Figure 29-5 highlights the implications of the various options. Not all adaptations are equally appropriate in all contexts. Understanding the baseline conditions and stresses (both climate and other) are important in understanding which climate change adaptation option will generate the greatest benefits. On small islands where resources are often limited, recognising the starting point for action is critical to maximising the benefits from adaptation. The following section considers the benefits of pursuing the various options. Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [INSERT FIGURE 29-5 HERE Figure 29-5: The impact of alternative climate change adaptation actions or policies.] 29.6.2.1. Building Adaptive Capacity with Traditional Knowledge, Technologies, and Skills on Small Islands As in previous IPCC Assessments, there is continuing strong support for the incorporation of indigenous knowledge into adaptation planning. However this is moderated by the recognition that current practices alone may not be adequate to cope with future climate extremes or trend changes The ability of a small island population to deal with current climate risks may be positively correlated with the ability to adapt to future climate change, but evidence confirming this remains limited (such as Lefale, 2010). Consequently, this section focuses on evidence for adaptive capacity that reduces vulnerability to existing stressors, enables adaptation to current stresses, and supports current disaster risk management. Traditional knowledge has proven to be useful in short term weather forecasting (e.g. Lefale, 2010) although evidence is inconclusive on local capacity to observe long-term climate change (e.g. Hornidge and Scholtes, 2011). In Solomon Islands, Lauer and Aswani (2010) found mixed ability to detect change in spatial cover of seagrass meadows. In Jamaica, Gamble et al. (2010) reported a high level of agreement between farmers perception of increasing drought incidence and statistical analysis of precipitation and vegetation data for the area. In this case farmers perceptions clearly validated the observational data and vice versa. Despite some claims that vulnerability reduction in indigenous communities in small islands may be best tackled by combining indigenous and Western knowledge in a culturally compatible and sustainable manner (Mercer et al., 2007), given the small number of studies in this area, there is not sufficient evidence to determine the effectiveness and limits to the use of traditional methods of weather forecasting under climate change on small islands. Traditional technologies and skills can be effective for current disaster risk management but there is currently a lack of supporting evidence to suggest that they will be equally appropriate under changing cultural conditions and future climate changes on islands. Campbell (2009) identified that traditional disaster reduction measures used in Pacific islands focused around maintaining food security, building community cooperation, and protecting settlements and inhabitants. Examples of actions to maintain food security include: the production and storage of food surpluses such as yam and breadfruit buried in leaf-lined pits to ferment; high levels of agricultural diversity to minimise specific damage to any one crop; and the growth of robust famine crops unused in times of plenty which could be used in emergencies (Campbell, 2009). Two discrete studies from Solomon Islands highlight the importance of traditional patterns of social organisation within communities to support food security under social and environmental change (Reenberg et al., 2008; Mertz et al., 2010). In both studies the strategy of relying on traditional systems of organization for farming and land use management have been shown to work effectively largely as there has been little cultural and demographic change. Nonetheless there are physical and cultural limits to traditional disaster risk management. In relation to the ability to store surplus production on atoll islands, on Rongelap in the Marshall Islands, surpluses are avoided, or are redistributed to support community bonds (Bridges and McClatchey, 2009). Further, traditional approaches that Pacific island communities have used for survival for millennia (such as building elevated settlements and resilient structures; and working collectively), have been abandoned or forgotten due to processes of globalisation, colonialism and development (Campbell, 2009). Ongoing processes of rapid urbanization, and loss of language and tradition suggest that traditional approaches may not always be efficacious in longer-term adaptation. Traditional construction methods have long been identified across the Pacific as a means of reducing vulnerability to tropical cyclones and floods in rural areas. In Solomon Islands traditional practices include: elevating concrete floors on Ontong Java to keep floors dry during heavy rainfall events; building low, aerodynamic houses with sago palm leaves as roofing material on Tikopia as preparedness for tropical cyclones; and in Bellona local perceptions are that houses constructed from modern materials and practices are more easily destroyed by tropical cyclones, implying that traditional construction methods are perceived to be more resilient in the face of extreme weather (Rasmussen et al., 2009). In parallel, Campbell (2009) documents the characteristics of traditional building styles (in Fiji, Samoa and Tonga) where relatively steep hipped roofs, well bound connections and joints, and airtight spaces with few windows or doors offer some degree of wind resistance. Traditional building measures can also reduce Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 damages associated with earthquakes as evidenced in Haiti (Audefroy, 2011). By reducing damage caused by other stresses (such as earthquakes), adaptive capacity is more likely to be maintained. The quality of home construction is critical to its wind-resistance. If inadequately detailed, home construction will fail irrespective of method. While some traditional measures could be challenged as potentially risky for example using palm leaves, rather than metal roofs as a preparation for tropical cyclone impacts the documentation of traditional approaches, with an evaluation of their effectiveness remains urgently needed. Squatter settlements in urban areas, especially on steep hillsides in the Caribbean often use poor construction practices frequently driven by poverty and inadequate building code enforcement (Prevatt et al., 2010). Traditional systems appear less effective when multiple civilization-nature stresses are introduced. For example in Reunion and Mayotte, population growth, and consequent rises in land and house prices have led low-income families to settle closer to hazardous slopes that are prone to landslides and to river-banks which are prone to flooding (Le Masson and Kelman, 2011). Traditional belief systems can also limit adaptive capacity. Thus, for example in two Fijian villages, approximately half of survey respondents identified divine will as the cause of climate change (Lata and Nunn, 2012). These findings reinforce earlier studies in Tuvalu (Mortreux and Barnett, 2009), and more widely across the Pacific (Barnett and Campbell, 2010). The importance of taking into account local interests and traditional knowledge in adaptation in small islands is emphasised by Kelman and West (2009) and McNamara and Westoby (2011), yet evidence does not yet exist that reveals the limits to such knowledge, such as in the context of rapid socio-ecological change, or the impact of belief systems on adaptive capacity. While there is clear evidence that traditional knowledge networks, technologies and skills can be used effectively to support adaptation in certain contexts, the limits to these tools are not well understood. To date research in the Pacific and Caribbean dominates small island climate change work. More detailed studies on small islands in the central and western Indian Ocean, the Mediterranean and the central and eastern Atlantic would improve understanding on this topic. 29.6.2.2. Addressing Risks on Small Islands Relative to other areas, small islands are disproportionately affected by current hydro-meteorological extreme events, both in terms of the percentage of the population affected, and losses as a percentage of GDP (Anthoff et al., 2010; Table 29-5). Under climate change the risks of damage and associated losses are expected to continue to rise (Nicholls and Cazenave, 2010). Yet much of the existing literature on climate risk in small islands does not consider how to address high future risks, but instead focuses on managing present day risks through risk transfer, risk spreading or risk avoidance. Risk transfer is largely undertaken through insurance; risk spreading through access to and use of common property resources, livelihood diversification, or mutual support through networks (see 29.6.2.3); and risk avoidance through structural engineering measures or migration (see 29.6.2.4). [INSERT TABLE 29-5 HERE Table 29-5: Top ten countries in the Asia-Pacific region based on absolute and relative physical exposure to storms and impact on GDP (between 1998 and 2009).] Risk transfer through insurance markets has had limited uptake in small islands, as insurance markets do not function as effectively as they do in larger locations, in part due to a small demand for the insurance products (Heger et al., 2008). In the case of insurance for farmers, researchers found that a lack of demand for insurance products (in their study countries: Grenada, Jamaica, Fiji and Vanuatu) meant an undersupply of customized food insurance products, which in turn contributed to a lack of demand for insurance (Angelucci and Conforti, 2010). Alternatives exist such as index-based schemes that provide payouts based on the crossing of a physical threshold, e.g. when rainfall drops below a certain level, rather than on drought damage sustained (Linnerooth-Bayer and Mechler, 2009). The potential for index-based insurance for climate stressors on islands is under-researched and there remains limited evidence of the long-term effectiveness of index-based or pooled-risk insurance in supporting household level adaptation. Small island governments also face expensive climate risk insurance. The Caribbean Catastrophe Risk Insurance Facility (CCRIF) which has been operating since 2007 pools Caribbean-wide country-level risks into Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 a central, more diversified risk portfolio offering lower premiums for participating national governments (CCRIF, 2008). The potential for a similar scheme in the Pacific is being explored (ADB, 2009; Cummins and Mahul, 2009). Risk can be spread socially e.g. through social networks and familial ties (see also 29.6.2.3), or ecologically, e.g. by changing resource management approach. Social networks can be used to spread risk among households. In Fiji, after Tropical Cyclone Ami in 2003, households whose homes were not affected by the cyclone increased their fishing effort to support those whose homes were damaged (Takasaki, 2011) mutual support formed a central pillar for community-based adaptation. In the case of natural systems, risks can be spread through enhancing representation of habitat types and replication of species e.g. through the creation of marine protected areas, around key refuges that protect a diversity of habitat, that cover an adequate proportion of the habitat and that protect critical areas such as nursery grounds and fish spawning aggregation areas (McLeod et al., 2009). Locally Managed Marine Areas which involve the local community in the management and protection of their local marine environment have proven to be effective in increasing biodiversity, and in reducing poverty in areas dependent on marine resources in several Pacific islands (Techera, 2008; Game et al., 2011). By creating a network of protected areas supported by local communities the risks associated with some forms of climate change can be spread and potentially reduced (Mills et al., 2010) although such initiatives may not preserve thermally sensitive corals in the face of rising SST. Risk avoidance through engineered structures can reduce risk from some climate-related hazards (medium evidence, medium agreement). In Jamaica, recommendations to reduce rainfall-driven land surface movements resulting in landslides include: engineering structures such as soil nailing, gabion baskets (i.e. cages filled with rocks), rip rapped surfaces (i.e. permanent cover with rock) and retaining walls together with engineered drainage systems (Miller et al., 2009b). Engineering principles to reduce residential damage from hurricanes have been identified, tested, and recommended for decades in the Caribbean. However, expected levels of success have often not been achieved due to inadequate training of construction workers, minimal inspection of new buildings, and lack of enforcement of building code requirements (Prevatt et al., 2010). Some island states do not even have the technical or financial capacity to build effective shore protection structures as highlighted by a recent assessment in south Tarawa, Kiribati (Duvat, 2013). In addition not all engineered structures are seen as effective risk avoidance mechanisms. In the Azores archipelago, a proliferation of permanent engineered structures along the coastline to prevent erosion have resulted in a loss of natural shoreline protection against wave erosion (Calado et al., 2011)). In Barbados it is recognized that seawalls can protect human assets in areas prone to high levels of erosion, however they can also cause sediment starvation in other areas, interfere with natural processes of habitat migration and cause coastal squeeze which may render them less desirable for long term adaptation (Mycoo and Chadwick, 2012) (see also 5.4.2.1). To reduce erosion risk an approach with less detrimental downstream effects that also supports tourism is beach nourishment. This is increasingly being recommended, for example in the Caribbean (Mycoo and Chadwick, 2012), the Mediterranean (Anagnostou et al., 2011), and western Indian Ocean (Duvat, 2009). Beach nourishment however is not without its challenges, as requirements such as site-specific oceanographic and wave climate data, adequate sand resources and critical engineering design skills may not be readily available in some small islands. 29.6.2.3. Working Collectively to Address Climate Impacts on Small Islands More attention is being focused on the relevance and application of community-based adaptation (CBA) principles to island communities, to facilitate adaptation planning and implementation (Warrick, 2009; Kelman et al., 2011) and to tackle rural poverty in resource dependent communities (Techera, 2008). CBA research is focusing on empowerment that helps people to help themselves e.g. through marine catch monitoring (Breckwoldt and Seidel, 2012), while addressing local priorities and building on local knowledge and capacity. This approach to adaptation is being promoted as an appropriate strategy for small islands, since it is something done with rather than to communities (Warrick, 2009). Nonetheless externally driven programs to encourage community-level action have produced some evidence of effective adaptation. Both Limalevu et al. (2010) and Dumaru (2010) describe the outcomes of externally-led pilot CBA projects (addressing water security and coastal management) implemented in villages across Fiji, notably: more effective management of local water resources through capacity building; enhanced knowledge of climate change; and, the establishment of mechanisms to facilitate greater access to Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 technical and financial resources from outside the community. More long term monitoring and evaluation of the effectiveness of community level action is needed. Collaboration between stakeholders can lessen the occurrence of simple mistakes that can reduce the effectiveness of adaptation actions (medium evidence, medium agreement). Evidence from the Eastern Caribbean suggests that adaptations taken by individual households to reduce landslide risk building simple retaining walls can be ineffective compared to community level responses (Anderson et al., 2011). Landslide risk can be significantly reduced through better hillside drainage. In the Eastern Caribbean, community groups, with input from engineers, have constructed these networks of drains to capture surface runoff, household roof-water and grey water. Case studies from Fiji and Samoa in which multi-stakeholder and multi-sector participatory approaches were used to help enhance resilience of local residents to the adverse impacts of disasters and climate change (Gero et al., 2011) further support this view. In the case of community based disaster risk reduction (CBDRR), Pelling (2011) notes that buy-in from local and municipal governments is needed, as well as strong pre-existing relationships founded on routine daily activities, to make CBDRR effective. Research from both Solomon Islands and the Cayman Islands reinforce the conclusion that drivers of community resilience to hazard maps closely onto factors driving successful governance of the commons, that is: community cohesion; effective leadership; and, community buy-in to collective action (Schwarz et al., 2011; Tompkins et al., 2008). Where community organisations are operating in isolation, or where there is limited coordination and collaboration community vulnerability is expected to increase (Ferdinand et al., 2012). Strong local networks, and trusting relationships between communities and government appear to be key elements in adaptation, in terms of maintaining sustainable agriculture and in disaster risk management (medium evidence, high agreement). All of these studies reinforce the earlier work of Barnett (2001), providing empirical evidence that supporting community-led approaches to disaster risk reduction and hazard management may contribute to greater community engagement with anticipatory adaptation. However, it is not yet possible to identify the extent to which climate resilience is either a coincidental benefit of island lifestyle and culture, or a purposeful approach, such as the community benefits gained from reciprocity among kinship groups (Campbell, 2009). 29.6.2.4. Addressing Long-Term Climate Impacts and Migration on Small Islands Sea-level rise poses one of the most widely recognized climate change threats to low-lying coastal areas on islands (29.3.1). However long term climate impacts depend on the type of island (see Figure 29-1) and the adaptation strategy adopted. Small island states have 16% of their land area in low elevation coastal areas (<10m) as opposed to a global average of 2%, and the largest proportion of low elevation coastal urban land area: 13% (along with Australia and New Zealand), in contrast to the global average of 8% (McGranahan et al., 2007). Statistics like these underpin the widely held view about small islands being overwhelmed by rising seas associated with sea-level rise (Loughry and McAdam, 2008; Yamamoto and Esteban, 2010; Gordon-Clark, 2012; Berringer, 2012; Dema, 2012; Lazrus, 2012; Laczko and Aghazarm, 2009). Yet there remains limited evidence as to which regions (Caribbean, Pacific, Indian Ocean, west African islands) will experience the largest sea-level rise (Willis and Church, 2012) and which islands will experience the worst climate impacts. Nicholls et al. (2011) have modeled impacts of 4°C warming, producing a 0.5 to 2.0m sea-level rise, to assess the impacts on land loss and migration. With no adaptation occurring, they estimate that this could produce displacement of between 1.2 and 2.2 million people from the Caribbean, Indian Ocean and Pacific Ocean. More research is needed to produce robust agreement on the impact of sea-level rise on small islands, and on the range of adaptation strategies that could be appropriate for different island types under those scenarios. Research into the possible un-inhabitability of islands has to be undertaken sensitively to avoid short-term risks (i.e. to avoid depopulation and ultimately island abandonment) associated with a loss of confidence in an islands future (McNamara and Gibson, 2009; McLeman, 2011). Due to the high costs of adapting on islands it has been suggested that there will be a need for migration (Nicholls et al., 2011; Gemenne, 2011; Biermann and Boas, 2010; Voccia 2012). Relocation and displacement are frequently cited as outcomes of sea-level rise, salinisation and land loss on islands (Byravan and Rajan, 2006; Kolmannskog and Trebbi, 2010; see also 29.3.3.3). Climate stress is occurring at the same time as the growth in rural to urban migration. The latter is leading to squatter settlements that strain urban infrastructure notably: sewerage, waste Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 management, transport and electricity (Jones, 2005; Connell and Lea, 2002). Urban squatters on islands often live in highly exposed locations, lacking basic amenities, leaving them highly vulnerable to climate risks (Baker, 2012). However, a lack of research in this area makes it difficult to draw clear conclusions on the impact of climate change on the growing number of urban migrants in islands. Recent examples of environmental stress driven relocation and displacement provide contemporary analogues of climate-induced migration. Evidence of post-natural disaster migration has been documented in the Caribbean in relation to hurricanes (McLeman and Hunter, 2010), and in the Carteret Islands, Papua New Guinea where during an exceptionally high inundation event in 2008 (see 29.5.1.1) islanders sought refuge on neighbouring Bougainville island (Jarvis, 2010). Drawing any strong conclusions from this literature is challenging as there is little understanding of how to measure the effect of the environmental signal in migration patterns (Afifi et al., 2013; Krishnamurthy, 2012). While the example of the Carteret Islands cannot be described as evidence of adaptation to climate change, it suggests that under some extreme scenarios island communities may need to consider relocating in the future (Gemenne, 2011). In reality, financial and legal barriers are expected to inhibit significant levels of international environmentally induced migration in the Pacific (Barnett and Chamberlain, 2010). 29.6.3. Barriers and Limits to Adaptation in Small Island Settings Since publication of the IPCC SAR in 1996, significant barriers to climate change adaptation strategies in island settings have been discussed in considerable detail. Barriers include inadequate access to financial, technological and human resources, issues related to cultural and social acceptability of measures, constraints imposed by the existing political and legal framework, the emphasis on island development as opposed to sustainability, and a tendency to focus on addressing short term climate variability rather than long term climate change, and community preferences for hard adaptation measures such as seawalls instead of soft measures such as beach nourishment (Sovacool, 2012). Heger et al. (2008) recognised that more diversified economies have more robust responses to climate stress, yet most small islands lack economies of scale in production, thus specialising in niche markets and developing monocultures (e.g. sugar or bananas). Non-sovereign island states face additional exogenous barriers to adaptation. For example, islands like Réunion and Mayotte benefit from the provision of social services somewhat similar to what obtains in the Metropole, but not the level of enforcement of building codes and land use planning as in France (Le Masson and Kelman, 2011). Owing to their nature and complexity, these constraints will not be easily eliminated in the short term and will require on-going attention if their impact is to be minimized over time. Exogenous factors, such as the comparatively few assessments of social vulnerability to climate change, adaptation potential or resilience for island communities (Barnett, 2010) limit current understanding. In part this is due to the particularities of islands both their heterogeneity and their difference from mainland locations as well as the limitations of climate models in delivering robust science for small islands. It remains the case that thirteen years after Nurse et al. (2001) noted that downscaled global climate models do not provide a complete or necessarily accurate picture of climate vulnerabilities on islands, there is still little climate impacts research that reflects local concerns and contexts (Barnett et al., 2008). While lack of access to adequate financial, technological and human resources is often cited as the most critical constraint, experience has shown that endogenous factors such as culture, ethics, knowledge and attitudes to risk are important in constraining adaptation. Translating the word climate into Marshallese implies cosmos, nature and culture as well as weather and climate (Rudiak-Gould, 2012). Such cultural misunderstandings can create both barriers to action and novel ways of engaging with climate change. The lack of local support (due to encroachment on traditional lands) for the development of new infiltration galleries to augment freshwater supply on Tarawa atoll, Kiribati, highlights the importance of social acceptability (Moglia et al., 2008a, 2008b). Such considerations have led to the conclusion that there is still much to be learned about the drivers of past adaptation and how mainstreaming into national programs and policies, widely acclaimed to be a virtually indispensable strategy, can practically be achieved (Mercer et al., 2007; Adger et al., 2009; Mertz et al., 2009). Notwithstanding the extensive and ever-growing body of literature on the subject, there is still a relatively low level of awareness and understanding at the community level on many islands about the nature of the threat posed by climate change (Nunn, 2009). Even where the threat has been identified, it is often not considered an urgent issue, or Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 a local priority, as exemplified in Malta (Akerlof et al., 2010) and Funafuti, Tuvalu (Mortreux and Barnett, 2009). Lack of awareness, knowledge and understanding can function as an effective barrier to the implementation and ultimate success of adaptation programs. This is borne out in both Fiji and Kiribati where researchers found that spiritual beliefs, traditional governance mechanisms, and a short term approach to planning were barriers to community engagement and understanding of climate change (Kuruppu, 2009; Lata and Nunn, 2012). Although widely acknowledged to be critical in small islands, few initiatives pay little more than perfunctory attention to the importance of awareness, knowledge and understanding in climate change adaptation planning. Hence, the renewed call for adaptation initiatives to include and focus directly on these elements on an ongoing basis (e.g., Crump, 2008; Kelman and West, 2009; Kelman, 2010; Kuruppu and Liverman, 2011; Gero et al., 2011) is timely, if these barriers are to be eventually removed. 29.6.4. Mainstreaming and Integrating Climate Change into Development Plans and Policies There is a growing body of literature that discusses the benefits and possibilities of mainstreaming or integrating climate change policies in development plans. Various mechanisms through which development agencies as well as donor and recipient countries can seek to capitalize on the opportunities to mainstream are beginning to emerge (see for example Klein et al., 2007; Mertz et al., 2009). Agrawala and van Aalst (2008) provide examples from Fiji and elsewhere, of where synergies (and trade-offs) can be found in integrating adaptation to climate change into development cooperation activities, notably in the areas of: disaster risk reduction, community-based approaches to development, and building adaptive capacity. Boyd et al. (2009) support the need for more rapid integration of adaptation into development planning, to ensure that adaptation is not side-lined, or treated separately from sectoral policies. Although there are synergies and benefits to be derived from the integration of climate change and development policies, care is needed to avoid institutional overlaps, and differences in language and approach which can give rise to conflict (Schipper and Pelling, 2006). Overall, there appears to be an emerging consensus around the views expressed by Swart and Raes (2007) that climate change and development strategies should be considered as complementary, and that some elements such as land and water management and urban, peri-urban and rural planning provide important adaptation, development and mitigation opportunities. While the potential to deliver such an integrated approach may be reasonably strong in urban centres on islands, there appears to be limited capacity to mainstream climate change adaptation into local decision making in out-lying islands or peripheral areas (Nunn et al., 2013). 29.7. Adaptation and Mitigation Interactions Greenhouse gas emissions from most small islands are negligible in relation to global emissions, yet small islands will most probably be highly impacted by climate change (Srinivasan, 2010). However, many small island governments and communities have chosen to attempt to reduce their greenhouse gas emissions because of the cost andthe potential co-benefits and synergies. Malta and Cyprus are obliged to do so in line with EU climate and energy policies. This section considers some of the inter-linkages between adaptation and mitigation on small islands and the potential synergies, conflicts, trade-offs and risks. Unfortunately there is relatively little research on the emissions reduction potential of small islands, and far less on the inter-linkages between climate change adaptation and emissions reduction in small islands. Therefore in this section a number of assumptions are made about how and where adaptation and mitigation actions interact. 29.7.1. Assumptions/Uncertainties Associated with Adaptation and Mitigation Responses Small islands are not homogeneous. Rather they have diverse geo-physical characteristics and economic structures (see Table 29-2, Figure 29-1). Following Nunn (2009) the combination of island geography and economic types informs the extent to which adaptation and mitigation actions might interact. The geography and location of islands affect their sensitivity to hydro-meteorological and related hazards such as cyclones, floods, droughts, invasive alien species, vector borne disease, and landslides. On the other hand the capacity of island residents to cope is often related to income levels, resources endowment, technology and knowledge (see 29.6.2). Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The potential for mitigation and emissions reductions in islands depends to a large extent on their size and stage of economic development. In the small and less developed islands key mitigation sectors including energy, transport, industry, built environment, agriculture, forestry, or waste management sectors are generally relatively small (Metz et al., 2007; Swart and Raes, 2007). Hence opportunities for emissions reductions are usually quite limited and are mostly associated with electricity generation and utilization of vehicles. More mitigation opportunities should exist in more economically advanced and larger islands that rely on forms of production that utilize fossil fuels, including manufacturing, and where vehicle usage is extensive and electricity driven home appliances, such as air conditioners and water heaters, are extensively used. In the absence of significant mitigation efforts at the global scale, adaptation interventions could become very costly and difficult to implement, once certain thresholds of change are reached (Nelson, 2011; Birkmann, 2011). Nicholls et al. (2011) make a similar observation with respect to coastal protection as a response to sea-level rise. They suggest that if global mean temperatures increase by around 4 0C (which may lead to sea-level rise between 0.5 m and 2 m) the likelihood of successful coastal protection in some locations such as low-lying small islands, will be low. Consequently, it is argued that the relocation of communities would be a likely outcome in such circumstances (Nicholls et al., 2011). 29.7.2. Potential Synergies and Conflicts Metz et al. (2007) suggest that adaptation and mitigation interactions occur in one of four main ways: adaptations that result in greenhouse emissions reduction; mitigation options that facilitate adaptation; policy decisions that couple adaptation and mitigation effects; and, trade-offs and synergies between adaptation and mitigation. Each of these opportunities is considered using three examples: coastal forestry, energy supply, and tourism. Small islands have relatively large coastal zones (in comparison to land area) and most development (as well as potential mitigation and adaptation activities) are located in the coastal zone. Coastal ecosystems (coral reefs, sea grasses and mangroves) play an important role in protecting coastal communities from wave erosion, tropical cyclones, storm surges, and even moderate tsunami waves (Cochard et al., 2008). Whilst coastal forests including both endemic and exotic species especially mangroves are seen as effective adaptation options ( bioshields Feagin et al., 2010) in the coastal zones, they also play an important role in mitigation as carbon sinks (van der Werf et al., 2009). Thus, the management and conservation of mangrove forests has the potential to generate synergies between climate change adaptation and mitigation. However, despite this knowledge population, development and agricultural pressures have constrained the expansion of island forest carbon stocks (Fox et al., 2010) while Gilman et al. (2008) note that such pressures can also reduce the buffering capacity of coastal vegetation systems. Renewable energy resources on small islands have only recently been considered within the context of long-term energy security (Praene et al., 2012; Chen et al., 2007). Stuart (2006) speculates that the lack of uptake of renewable technologies to date might be due to historical commitments to conventional fossil fuel-based infrastructure, and a lack of resources to undertake research and development of alternatives. Those islands that have introduced renewable energy technologies have often done so with support from international development agencies (Dornan, 2011). Despite this, there remain significant barriers to the wider institutionalization of renewable technologies in small islands. Research in Europe and the United States has shown the mitigation and cost savings benefits of Energy Service Companies (ESCOs): companies that enter into medium-to-long term performance-based contracts with energy users, invest in energy efficiency measures in buildings and firms, and profit from the ensuing energy savings measures for the premises (see for example Steinberger et al., 2009). Potential benefits exist in creating the opportunity for ESCOs to operate in small islands. Preliminary evidence from Fiji suggests that if the incentive mechanisms can be resolved, and information asymmetries between service providers and users can be aligned, ESCOs could provide an opportunity to expand renewable technologies (Dornan, 2009). IPCC (2011) presents examples of opportunities for renewable energy, including wind energy sources, as deployed in the Canary Islands. The transition towards renewable energy sources away from fossil fuel dependence has been partly driven by economic motives, notably to avoid oil price volatility and its impact. The development of hydro-power (in Fiji for Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 example) necessitates protection and management of the water catchment zones, and thus could lead to improved management of the water resources a critical adaptation consideration for areas expected to experience a decrease in average rainfall as a result of climate change. Whilst the cost effectiveness of renewable technologies is critical, placing it within the context of water adaptation could enhance project viability (Dornan, 2009). Cost-benefit analyses have shown that in southeast Mediterranean islands photovoltaic generation and storage systems may be more cost-effective than existing thermal power stations (Kaldellis, 2008; Kaldellis et al., 2009). Energy prices in small islands are among the highest anywhere in the world, mainly due to their dependence on imported fossil fuel, and limited ability to reap the benefits of economies of scale including bulk buying. Recent studies show that the energy sectors in small islands may be transformed into sustainable growth entities mainly through the judicious exploitation of renewable energy sources, combined with the implementation of energy efficiency measures (van Alphen et al., 2008; Banuri, 2009; Mohanty, 2012; Rogers et al., 2012). Realising the potential for such transformation, the countries comprising the Alliance of Small Island States (AOSIS) launched SIDS Dock, which is intended to function as a docking station to connect the energy sector in SIDS with the international finance, technology and carbon markets with the objective of pooling and optimizing energy efficiency goods and services for the benefit of the group. This initiative seeks to decrease energy dependence in SIDS, while generating financial resources to support low carbon growth and adaptation interventions. Many small islands rely heavily on the foreign exchange from tourism to expand and develop their economies, including the costs of mitigation and adaptation. Tourism, particularly in small islands, often relies on coastal and terrestrial ecosystems to provide visitor attractions and accommodation space. Recognising the relationship between ecosystem services and tourism in Jamaica, Thomas-Hope and Jardine-Comrie (2007) suggest that sustainable tourism planning should include activities undertaken by the industry, that is tertiary treatment of waste, and re-use of water, as well as composting organic material and investing in renewable energy. Gössling and Schumacher (2010) and others who have examined the linkages between greenhouse gas emissions and sustainable tourism argue that the tourism sector (operators and tourists) should pay to promote sustainable tourism, especially where they benefit directly from environmental services sustained by these investments. 29.8. Facilitating Adaptation and Avoiding Maladaptation While there is a clear consensus that adaptation to the risks posed by global climate change is necessary and urgent in small islands, the implementation of specific strategies and options is a complex process that requires critical evaluation of multiple factors, if expected outcomes are to be achieved (Kelman and West, 2009; Barnett and O Neill, 2012). These considerations may include, inter alia, prior experience with similar or related threats, efficacy of the strategies or options and their co-benefits, costs (monetary and non-monetary), availability of alternatives and social acceptability. In addition, previous work (e.g. Adger et al., 2005) has emphasised the relevance of scale as a critical factor when assessing the efficacy and value of adaptation strategies, as the extent to which an option is perceived to be a success, failure or maladaptive may be conditioned by whether it is being assessed as a response to climate variability (shorter-term) or climate change (longer-term). As in other regions, adaptation in islands is locally delivered and context specific (Tompkins et al., 2010). Yet, sectors and communities on small islands are often so intricately linked that there are many potential pathways that may lead to maladaptation, be it via increased greenhouse gas emissions, foreclosure of future options, or burdensome opportunity costs on local communities. There is also a concern that some types of interventions may actually be maladaptive. For example, Barnett and O Neill (2012) suggest that strategies such as resettlement and migration should be regarded as options of last resort on islands, as they may actually discourage viable adaptation initiatives, by fostering over-dependence on external support. They further argue that a priori acceptance of adaptation as an efficacious option for places like the Pacific islands, may also act as a disincentive for reducing greenhouse gas emissions (Barnett and O Neill, 2012). Notwithstanding the observations of Barnett and O Neill (2012), there is a concern that early foreclosure of this option might well prove maladaptive, if location-specific circumstances show such action to be efficacious in the longer-term. For example, Bunce et al. (2009) have shown that as an adaptive response to poverty, young fishers Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 from Rodrigues Island periodically resort to temporary migration to the main capital island, Mauritius, where greater employment prospects exist. The case study of the residents of Nauru who contemplated resettlement in Australia after the collapse of phosphate mining (their only revenue source) in the1950s, provides helpful insight about the complex social, economic and cultural challenges associated with environmentally triggered migration (Tabucanon and Opeskin, 2011). Negotiations with the Government of Australia collapsed before a mutually acceptable agreement was reached, and the Nauruans opted to abandon the proposal to relocate (Tabucanon and Opeskin, 2011). Overall however, it is suggested that states contemplating long term, off-island migration may wish to consider early proactive planning, as resettlement of entire communities might prove to be socially, culturally and economically disruptive (Campbell, 2010; McMichael et al., 2012; refer also to 29.3.3.3). A related challenge facing small islands is the need to find the middle ground between resettlement and objective assessment of other appropriate adaptation choices. Similarly, while insurance is being promoted as an element of the overall climate change response strategy in some island regions, e.g. the Caribbean, concerns have been expressed about possible linkages to maladaptation. The potential consequences include the imposition of exorbitant premiums that are beyond the capacity of resource- scarce governments as the perception of climate change risks increase, discriminatory coverage of sectors that may not align with local priorities, and tacit encouragement for the state, individuals and the private sector to engage in behavior that is not risk-averse, e.g. development in hazard-prone areas (Herweijer et al., 2009; Linnerooth-Bayer et al., 2011; van Nostrand and Nevius, 2011; Thomas and Leichenko, 2011). Likewise, although the exploitation of renewable energy is vital to the sustainable development of small islands, more attention needs to be paid to the development of energy storage technologies, if rapid transition from conventional fuels is to be achieved in an efficient manner. This is especially important in the case of intermittent energy sources (e.g. solar and wind), as the cost of current storage technologies can frustrate achievement of full conversion to renewable energy. Thus to avoid the possibility of maladaptation in the sector, countries may wish to consider engaging in comprehensive planning, including considerations relating to energy storage (Krajaèiæ et al., 2010; Bazilian et al., 2011). Recent studies have demonstrated that opportunities exist in island environments for avoiding maladaptation. Studies have shown for example that decisions about adaptation choices and their implementation are best facilitated where there is constructive engagement with the communities at risk, in a manner that fosters transparency and trust (López-Marrero, 2010; van Aalst et al., 2008). Further, some analysts argue that adaptation choices are often subjective in nature and suggest that participatory stakeholder involvement can yield valuable information about the priorities and expectations that communities attach to the sector for which adaptation is being sought. The point is underscored by Moreno and Becken (2009) whose study of the tourism sector on the Mamanuca islands (Fiji), which clearly demonstrates that approaches which explicitly integrate stakeholders into each step of the process from vulnerability assessment right through to consideration of alternatives measures can provide a sound basis for assisting destinations with the implementation of appropriate adaptation interventions. This view is supported by Dulal et al. (2009), who argue that the most vulnerable groups in the Caribbean - the poor, elderly, indigenous communities and rural children will be at greater risk of being marginalized, if adaptation is not informed by equitable and participatory frameworks. Other studies reveal that new paradigms whose adoption can reduce the risk of maladaptation in island environments, are emerging across various sectors. In the area of natural resource management, Hansen et al. (2010) suggest that the use of protected areas for climate refugia, reduction of non-climate stressors on ecosystems, adoption of adaptive management approaches combined with reduction of greenhouse gas emissions wherever possible, may prove to be more effective response strategies than traditional conservation approaches. Other strategic approaches, including the implementation of multi-sectoral and cross-sectoral measures, also facilitate adaptation in a more equitable, integrated and sustainable manner. Similarly, no-regret measures such as wastewater recycling, trickle irrigation, conversion to non-fossil fuel based energy and transportation which offer collateral benefits with or without the threat of climate change, and low-regret strategies, which may only increase existing operational costs marginally, are becoming increasingly attractive options to island governments (Gravelle and Mimura, 2008; Heltberg et al., 2009; Howard et al., 2010). Together, these constitute valid risk management approaches, as they are designed to assist communities in making prudent, but necessary decisions in the face of an uncertain future. Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Some authors suggest that caution is needed to ensure that donors are not driving the adaptation and mitigation agenda in small islands, as there is a risk that donor-driven adaptation or mitigation may not always address the salient challenges on small islands, and may lead to inadequate adaptation or a waste of scarce resources (Barnett, 2010; Nunn, 2009). Others argue that donor-led initiatives may unintentionally cause enhanced vulnerability by supporting adaptation strategies that are externally derived, rather than optimizing the benefits of local practices that have proven to be efficacious through time (Reenberg et al., 2008; Kelman and West, 2009; Campbell and Beckford, 2009). 29.9. Research and Data Gaps Several advances have taken place in our understanding of the observed and potential effects of climate change on small islands since the AR4. These cover a range of themes including: dynamic downscaling of scenarios appropriate for small islands; impacts of trans-boundary processes generated well beyond the borders of an individual nation or island; barriers to adaptation in small islands and how they may be overcome; the relationships between climate change adaptation and disaster risk reduction; and, the relationships between climate change adaptation, maladaptation and sustainable development. It is also evident that much further work is required on these themes in small island situations, especially comparative research. Important information and data gaps and many uncertainties still exist on impacts, vulnerability and adaptation in small islands. These include: Lack of climate change and socio-economic scenarios and data at the required scale for small islands. Although some advances have been made (ABoM and CSIRO, 2011a, 2011b; Taylor et al., 2007), much of the work in the Caribbean, Pacific, Indian Ocean and Mediterranean islands, is focused at the regional scale rather than being country specific. Since most socio-economic decisions are taken at the local level, there is need for a more extensive database of simulations of future small island climates and socio-economic conditions at smaller spatial scales. Difficulties in detecting and attributing past impacts on small islands to climate change processes. Further investigation of the observed impacts of weather, climate and ocean events that may be related to climate change is required to clarify the relative role of climate change and non-climate change drivers. Uncertainty in the projections is not a sufficiently valid reason to postpone adaptation planning in small islands. In several small islands adaptation is being progressed without a full understanding of past or potential impacts and vulnerability. Whilst assessment of future impacts is hampered because of uncertainty in climate projections at the local island level, alternative scenarios based on a general understanding of broad trends could be used in vulnerability and sensitivity studies to guide adaptation strategies. Need for a range of climate change-related projections beyond temperature and sea-level. Generally climate-model projections of temperature and sea-level have been satisfactory, but there are strong requirements for projections for other variables that are of critical importance to small islands. These include rainfall and drought, wind direction and strength, tropical storms and wave climate, and recognition that trans-boundary processes are also significant in a small island context. While some such work has been undertaken for some parts of the Pacific (ABoM and CSIRO, 2011a, 2011b), similar work still needs to be carried out in other small island regions. In addition, the reliability of existing projections for some of the other parameters needs to be improved and the data should be in suitable formats for use in risk assessments. Need to acknowledge the heterogeneity and complexity of small island states and territories. Although small islands have several characteristics in common, neither the variety nor complexity of small islands is sufficiently reflected in the literature. Thus, transferring data and practices from a continental situation, or from one small island state to another, needs to be done with care and in a manner that takes full cognizance of such heterogeneity and complexity. Within country/territory differences need to be better understood. Many of the environmental and human impacts reported in the literature on islands have been attributed to the whole country, when in fact they refer only to the major centre or town or region. There is need for more work on rural areas, outer islands and secondary communities. Several examples of such research have been cited in this chapter. Also Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 it should be noted that some small island states are single islands and others highly fragmented multiple islands. Lack of investment and attention to climate and environmental monitoring frameworks in small islands. A fundamental gap in the ability to improve empirical understanding of present and future climate change impacts is the lack of climate and environmental monitoring frameworks that in turn hampers the level of confidence with which adaptation responses can be designed and implemented. Economic and social costs of climate change impacts and adaptation options are rarely known. In small island states and territories the costs of past weather, climate and ocean events is poorly known and further research is required to identify such costs, and to determine the economic and societal costs of climate change impacts and the costs of adaptation options to minimize those impacts. The foregoing list is a sample of the gaps, needs and research agenda that urgently need to be filled for small islands. While some countries have begun to fill these gaps, this work needs to be replicated and expanded across all island regions to improve the database available for ongoing climate change assessments. Such information would raise the level of confidence in the adaptation planning and implementation process in small islands. Frequently Asked Questions FAQ 29.1: Why is it difficult to detect and attribute changes on small islands to climate change? [to be inserted in Section 29.3.1.1] In the last two or three decades many small islands have undergone substantial changes in human settlement patterns and in socio-economic and environmental conditions. Those changes may have masked any clear evidence of the effects of climate change. For example, on many small islands coastal erosion has been widespread and has adversely affected important tourist facilities, settlements, utilities and infrastructure. But specific case studies from islands in the Pacific, Indian and Atlantic oceans and the Caribbean have shown that human impacts play an important role in this erosion, as do episodic extreme events that have long been part of the natural cycle of events affecting small islands. So while coastal erosion is consistent with models of sea-level rise resulting from climate change, determining just how much of this erosion might have been caused by climate change impacts is difficult. Given the range of natural processes and human activities that could impact the coasts of small islands in the future, without more and better empirical monitoring the role of climate change-related processes on small islands may continue to be difficult to identify and quantify. FAQ 29.2: Why is the cost of adaptation to climate change so high in small islands? [to be inserted after Section 29.3.3.4] Adaptation to climate change that involves infrastructural works generally require large up-front overhead costs, which in the case of small islands cannot be easily downscaled in proportion to the size of the population or territory. This is a major socio-economic reality that confronts many small islands, notwithstanding the benefits that could accrue to island communities through adaptation. Referred to as indivisibility in economics, the problem can be illustrated by the cost of shore protection works aimed at reducing the impact of sea-level rise. The unit cost of shoreline protection per capita in small islands is substantially higher than the unit cost for a similar structure in a larger territory with a larger population. This scale-reality applies throughout much of a small island economy including the indivisibility of public utilities, services and all forms of development. Moreover, the relative impact of an extreme event such as a tropical cyclone that can affect most of a small island s territory has a disproportionate impact on that state s GDP, compared to a larger country where an individual event generally affects a small proportion of its total territory and its GDP. The result is relatively higher adaptation and disaster risk reduction costs per capita in countries with small populations and areas, especially those that are also geographically isolated, have a poor resource base and high transport costs FAQ 29.3: Is it appropriate to transfer adaptation and mitigation strategies between and within small island countries and regions? [to be inserted after Section 29.7.2] While lessons learned from adaptation and mitigation experiences in one island or island region may offer some guidance, caution must be exercised to ensure that the transfer of such experiences is appropriate to local biophysical, social, economic, political, and cultural circumstances. If this approach is not purposefully incorporated Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 into the implementation process, it is possible that maladaptation and inappropriate mitigation may result. It is therefore necessary to carefully assess the risk profile of each individual island so as to ensure that any investments in adaptation and mitigation are context specific. The varying risk profiles between individual small islands and small island regions have not always been adequately acknowledged in the past. References ADB, 2009: Natural Catastrophe Risk Insurance Mechanisms for Asia and the Pacific. Asian Development Bank (ADB), Manila, Philippines. Adger, W.N. and K. Brown, 2009: Vulnerability and resilience to environmental change: Ecological and social perspectives. In: A Companion to Environmental Geography. [Castree, N., D. Demeritt, D. Liverman, and B. Rhoads(eds.)]. Blackwell Publishing Ltd, Oxford, UK, pp. 109-122. Adger, W., N. Arnell, E. Tompkins, 2005: Successful adaptation to climate change across scales. Global Environmental Change, 5(2), 77-86. Adger, W., S. Dessai, M. Goulden, M. Hulme, I. Lorenzoni, D. Nelson, L. Naess, J. Wolf, and A. Wreford, 2009: Are there social limits to adaptation to climate change? Climatic Change, 93(3), 335-354. Afifi, T., K. Warner, and T. Rosenfeld, 2013: Environmentally induced migration, theoretical and methodological research challenges. The Encyclopedia of Global Human Migration. Blackwell Publishing Ltd. Agrawala, S. and M. van Aalst, 2008: Adapting development cooperation to adapt to climate change. Climate Policy, 8(2), 183-193. Akerlof, K., R. DeBono, P. Berry, A. Leiserowitz, C. Roser-Renouf, K. Clarke, A. Rogaeva, M.C. Nisbet, M.R. Weathers, and E.W. Maibach, 2010: Public perceptions of climate change as a human health risk: surveys of the United States, Canada and Malta. International Journal of Environmental Research and Public Health, 7(6), 2559-2606. Albins, M.A. and M.A. Hixon, 2008: Invasive Indo-Pacific lionfish Pterois volitans reduce recruitment of Atlantic coral-reef fishes. Marine Ecology Progress Series, 367, 233-238. Albins, M.A. and M.A. Hixon, 2013: Worst case scenario: potential long-term effects of invasive predatory lionfish (Pterois volitans) on Atlantic and Caribbean coral-reef communities. Environmental Biology of Fishes, 96, 1151-1157. Aligizaki, K. and G. Nikolaidis, 2008: Morphological identification of two tropical dinoflagellates of the genera Gambierdiscus and Sinophysis in the Mediterranean Sea. Journal of Biological Research-Thessaloniki, 9, 75- 82. Al-Jeneid, S., M. Bahnassy, S. Nasr, and M. El Raey, 2008: Vulnerability assessment and adaptation to the impacts of sea level rise on the Kingdom of Bahrain. Mitigation and Adaptation Strategies for Global Change, 13(1), 87-104. Alling, A., O. Doherty, H. Logan, L. Feldman, and P. Dustan, 2007: Catastrophic coral mortality in the remote Central Pacific Ocean: Kiribati, Phoenix Islands. Atoll Research Bulletin, (551), 1-21. Anagnostou, C., P. Antoniou, and G. Hatiris, 2011: Erosion of a depositional coast in NE Rhodos island (SE Greece) and assessment of the best available measures for coast protection. Journal of Coastal Research, SI 64, 1316-1319. Anderson, M.G., E. Holcombe, J.R. Blake, F. Ghesquire, N. Holm-Nielsen, and T. Fisseha, 2011: Reducing landslide risk in communities: Evidence from the Eastern Caribbean. Applied Geography, 31(2), 590-599. Angelo, C.L. and C.C. Daehler, 2013: Upward expansion of fire-adapted grasses along a warming tropical elevation gradient. Ecography, 36, 551-559. Angelucci, F. and P. Conforti, 2010: Risk management and finance along value chains of Small Island Developing States. Evidence from the Caribbean and the Pacific. Food Policy, 35(6), 565-575. Anthoff, D., R. Nicholls, and R. Tol, 2010: The economic impact of substantial sea-level rise. Mitigation and Adaptation Strategies for Global Change, 15(4), 321-335. Atkinson, C.T. and D.A. LaPointe, 2009: Introduced avian diseases, climate change, and the future of Hawaiian honeycreepers. Journal of Avian Medicine and Surgery, 23(1), 53-63. Audefroy, J.F., 2011: Haiti: post-earthquake lessons learned from traditional construction. Environment and Urbanization, 23(2), 447-462. Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Australian Bureau of Meteorology and CSIRO, 2011a: Climate Change in the Pacific: Scientific Assessment and New Research. Volume 1: Regional Overview, pp 1-257. Australian Bureau of Meteorology and CSIRO, 2011b: Climate Change in the Pacific: Scientific Assessment and New Research. Volume 2: Country Reports Overview, pp 1-273. Baider, C. and F.B.V. Florens, 2011: Control of invasive alien weeds averts imminent plant extinction. Biological Invasions, 13(12), 2641-2646. Baker, J.L., 2012: Climate Change, Disaster Risk, and the Urban Poor: Cities Building Resilience for a Changing World. World Bank Publications, Washington, D.C., pp. 297. Ballu, V., M.N. Bouin, P. Simeoni, W.C. Crawford, S. Calmant, J.-. Bore, T. Kanas, and B. Pelletier, 2011: Comparing the role of absolute sea-level rise and vertical tectonic motions in coastal flooding, Torres Islands (Vanuatu). Proceedings of the National Academy of Sciences, 108(32), 13019-13022. Banuri, T., 2009: Climate change and sustainable development. Natural Resources Forum, 33(4), 257-258. Barnett, J., 2001: Adapting to climate change in Pacific Island Countries: The problem of uncertainty. World Development, 29(6), 977-993. Barnett, J., S. Lambert, and I. Fry, 2008: The hazards of indicators: insights from the environmental vulnerability index. Annals of the Association of American Geographers, 98(1), 102-119. Barnett, J. and J. Campbell, 2010: Climate Change and Small Islands States: Power, Knowledge and the South Pacific. Earthscan, London and Washington D.C., pp. 218. Barnett, J. and S.J. O'Neill, 2012: Islands, resettlement and adaptation. Nature Climate Change, 2, 8-10. Barnett, J. and N. Chamberlain, 2010: Migration as Climate Change Adaptation: Implications for the Pacific. In: Climate Change and Migration, South Pacific Perspectives. [BURSON, B. (ed.)]. Institute of Policy Studies, Victoria University of Wellington, Wellington, New Zealand, pp. 51-60. Barnett, J. and M. Webber, 2010: Accommodating migration to promote adaptation to climate change. Policy Research Working Paper No 5270. The World Bank, pp. 62. Barnett, J., 2011: Dangerous climate change in the Pacific Islands: food production and food security. Regional Environmental Change, 11(1), 229-237. Barnett, J., 2010: Climate change science and policy, as if people mattered. In: Climate change, ethics and human security. [O'Brien, K., S. Clair, A. Lera, and B. Kristoffersen(eds.)]. Cambridge University Press, pp. 47-62. Bassiouni, M. and D.S. Oki, 2013: Trends and shifts in streamflow in Hawai i, 1913 2008. Hydrological Processes, 27(10), 1484-1500. Bazilian, M., H. Rogner, M. Howells, S. Hermann, D. Arent, D. Gielen, P. Steduto, A. Mueller, P. Komor, R.S.J. Tol, and K.K. Yumkella, 2011: Considering the energy, water and food nexus: Towards an integrated modelling approach. Energy Policy, 39(12), 7896-7906. Becker, M., B. Meyssignac, C. Letetrel, W. Llovel, A. Cazenave, and T. Delcroix , 2012: Sea level variations at tropical Pacific islands since 1950. Global and Planetary Change, 80-81, 85-98. Bedford, R. and C. Bedford, 2010: International Migration and Climate Change: A Post-Copenhagen Perspective on Options for Kiribati and Tuvalu. In: Climate Change and Migration: South Pacific Perspectives. [Burson, B. (ed.)]. Institute of Policy Studies, pp. 89-134. Béguin, A., S. Hales, J. Rocklöv, C. Aström, V.R. Louis, and R. Sauerborn, 2011: The opposing effects of climate change and socio-economic development on the global distribution of malaria. Global Environmental Change, 21(4), 1209-1214. Bell, J.D., C. Reid, M.J. Batty, P. Lehodey, L. Rodwell, A.J. Hobday, J.E. Johnson, and A. Demmke, 2013: Effects of climate change on oceanic fisheries in the tropical Pacific: implications for economic development and food security. Climatic Change, 119(1), 199-212. Bell, J.D., J.E. Johnson, and A.J. Hobday (eds.), 2011: Vulnerability of tropical Pacific fisheries and aquaculture to climate change. Secretariat of the Pacific Community, Noumea, New Caledonia, pp. 925. Benning, T.L., D. LaPointe, C.T. Atkinson, and P.M. Vitousek, 2002: Interactions of climate change with biological invasions and land use in the Hawaiian Islands: Modeling the fate of endemic birds using a geographic information system. Proceedings of the National Academy of Sciences, 99, 14246-14249. Berringer, A.C., 2012: Climate Change and Emigration: Comparing Sinking Islands and Jamaica. Naturally Immigrants, January, 106-120. Biermann, F. and I. Boas, 2010: Preparing for a warmer world: Towards a global governance system to protect climate refugees. Global Environmental Politics, 10(1), 60-88. Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bijlsma, L., C.N. Ehler, R.J.T. Klein, S.M. Kulshrestha, R.F. McLean, N. Mimura, R.J. Nicholls, L.A. Nurse, H. Perez Nieto, E.Z. Stakhiv, R.K. Turner, and R.A. Warrick, 1996: Coastal zones and small islands. In: Climate Change 1995: Impacts, Adaptations and Mitigation of Climate Change: Scientific - Technical Analyses. Contribution of Working Group II to the Second Assessment Report of the Intergovernmental Panel on Climate Change. [Watson, R.T., M.C. Zinyowera, and R.H. Moss (eds.)]. Cambridge University Press, Cambridge, UK and New York, USA, pp. 289-324. Birkmann, J., 2011: First-and second-order adaptation to natural hazards and extreme events in the context of climate change. Natural Hazards, 58(2), 811-840. Bishop, M.L., 2012: The political economy of small states: enduring vulnerability? Review of International Political Economy, 19(5), 942-960. Black, R., W.N. Adger, N.W. Arnell, S. Dercon, A. Geddes, and D. Thomas, 2011: The effect of environmental change on human migration. Global Environmental Change, 21(Supplement 1), S3-S11. Blackburn, T.M., P. Cassey, R.P. Duncan, K.L. Evans, and K.J. Gaston, 2004: Avian extinction and mammalian introductions on oceanic islands. Science, 305, 1955-1958. Blancard, S. and J. Hoarau, 2013: A new sustainable human development indicator for small island developing states: A reappraisal from data envelopment analysis. Economic Modelling, 30, 623-635. Bokhorst, S., A. Huiskes, P. Convey, and R. Aerts, 2007: The effect of environmental change on vascular plant and cryptogam communities from the Falkland Islands and the Maritime Antarctic. BMC Ecology, 7(1), 15. Bokhorst, S., A. Huiskes, P. Convey, P.M. van Bodegom, and R. Aerts, 2008: Climate change effects on soil arthropod communities from the Falkland Islands and the Maritime Antarctic. Soil Biology and Biochemistry, 40(7), 1547-1556. Boyd, E., N. Grist, S. Juhola, and V. Nelson, 2009: Exploring development futures in a changing climate: Frontiers for development policy and practice. Development Policy Review, 27(6), 659-674. Brandt, M.E. and J.W. McManus, 2009: Disease incidence is related to bleaching extent in reef-building corals. Ecology, 90(10), 2859-2867. Breckwoldt, A. and H. Seidel, 2012: The need to know what to manage community-based marine resource monitoring in Fiji. Current Opinion in Environmental Sustainability, 4(3), 331-337. Bridges, K.W. and W.C. McClatchey, 2009: Living on the margin: Ethnoecological insights from Marshall Islanders at Rongelap atoll. Global Environmental Change, 19(2), 140-146. Briguglio, L., G. Cordina, N. Farrugia, and S. Vella, 2009: Economic vulnerability and resilience: Concepts and measurements. Oxford Development Studies, 37(3), 229-247. Brown, J.N., J.R. Brown, C. Langlais, R. Colman, J. Risbey, B.F. Murphy, A. Moise, A.S. Gupta, I. Smith, and L. Wilson, 2013a: Exploring qualitative regional climate projections: a case study for Nauru. Climate Research, DOI 10.3354/cr01190 (http://www.int-res.com/prepress/c01190.html). Brown, J.N., A.S. Gupta, J.R. Brown, L.C. Muir, J.S. Risbey, P. Whetton, X. Zhang, A. Ganachaud, B. Murphy, and S.E. Wijffels, 2013b: Implications of CMIP3 model biases and uncertainties for climate projections in the western tropical Pacific. Climatic Change, 119(1), 147-161. Bunce, M., L. Mee, L.D. Rodwell, and R. Gibb, 2009: Collapse and recovery in a remote small island A tale of adaptive cycles or downward spirals? Global Environmental Change, 19(2), 213-226. Bush, D.M., W.J. Neal, and C.W. Jackson, 2009: Summary of Puerto Rico's vulnerability to coastal hazards: Risk, mitigation, and management with examples. Geological Society of America Special Papers, 460, 149-165. Buzinde, C.N., D. Manuel-Navarrete, E.E. Yoo, and D. Morais, 2010: Tourists Perceptions in a Climate of Change: Eroding Destinations. Annals of Tourism Research, 37(2), 333-354. Byravan, S. and S.C. Rajan, 2006: Providing new homes for climate change exiles. Climate Policy, 6(2), 247-252. Cahoon, D.R., P. Hensel, J. Rybczyk, K.L. McKee, C.E. Proffitt, and B.C. Perez, 2003: Mass tree mortality leads to mangrove peat collapse at Bay Islands, Honduras after Hurricane Mitch. Journal of Ecology, 91(6), 1093-1105. Calado, H., P. Borges, M. Phillips, K. Ng, and F. Alves, 2011: The Azores archipelago, Portugal: improved understanding of small island coastal hazards and mitigation measures. Natural Hazards, 58(1), 427-444. Cambers, G., 2009: Caribbean beach changes and climate change adaptation. Aquatic Ecosystem Health and Management, 12(2), 168-176. Campbell, D. and C. Beckford, 2009: Negotiating Uncertainty: Jamaican Small Farmers Adaptation and Coping Strategies, Before and After Hurricanes A Case Study of Hurricane Dean. Sustainability, 1(4), 1366-1387. Campbell, J.D., M.A. Taylor, T.S. Stephenson, R.A. Watson, and F.S. Whyte, 2011: Future climate of the Caribbean from a regional climate model. International Journal of Climatology, 31(12), 1866-1878. Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Campbell, J., 2009: Islandness: Vulnerability and resilience in Oceania. Shima: The International Journal of Research into Island Cultures, 3, 85-97. Campbell, J., 2010: Climate Change and Population Movement in Pacific Island Countries. In: Climate Change and Migration. [Burson, B. (ed.)]. Institute of Policy Studies, pp. 29-50. Campbell, J.R., M. Goldsmith, and K. Koshy, 2005: Community relocation as an option for adaptation to the effects of climate change and climate variability in Pacific Island countries (PICs). Asia-Pacific Network for Global Change Research, pp. 61. Campbell, S.J., L.J. McKenzie, and S.P. Kerville, 2006: Photosynthetic responses of seven tropical seagrasses to elevated seawater temperature. Journal of Experimental Marine Biology and Ecology, 330(2), 455-468. Cantin, N.E., A.N. Cohen, K.B. Karnauskas, A.N. Tarrant, and D.C. McCorkle, 2010: Ocean Warming slows coral growth in the Central Red Sea. Science, 329(5989), 322-325. Cardona, O.D., 2007: Indicators of disaster risk and risk management. Inter-American Development Bank, Washington D.C. Carilli, J.E., R.D. Norris, B. Black, S.M. Walsh, and M. McField, 2010: Century-scale records of coral growth rates indicate that local stressors reduce coral thermal tolerance threshold. Global Change Biology, 16(4), 1247-1257. Carpenter, R.C. and P.J. Edmunds, 2006: Local and regional scale recovery of Diadema promotes recruitment of scleractinian corals. Ecology Letters, 9(3), 268-277. Cashman, A., L. Nurse, and J. Charlery, 2010: Climate change in the Caribbean: The water management implications. Journal of Environment and Development, 19(1), 42-67. Caujape-Castells, J., A. Tye, D.J. Crawford, A. Santos-Guerra, A. Sakai, K. Beaver, W. Lobin, F. Vincent Florens, M. Moura, and R. Jardim, 2010: Conservation of oceanic island floras: present and future global challenges. Perspectives in Plant Ecology, Evolution and Systematics, 12(2), 107-129. Cazenave, A. and W. Llovel, 2010: Contemporary Sea Level Rise. Annual Review of Marine Science, 2, 145-173. Cazenave, A. and F. Remy, 2011: Sea level and climate: measurements and causes of changes. Wiley Interdisciplinary Reviews: Climate Change, 2(5), 647-662. CCRIF, 2008: The Caribbean Catastrophe Risk Insurance Facility Annual Report 2007-2008. CCSP, 2008: Weather and Climate Extremes in a Changing Climate. Regions of Focus: North America, Hawaii, Caribbean, and U.S. Pacific Islands. A Report by the U.S. Climate Change Science Program and the Subcommittee on Global Change Research [K.R. Thomas, G.A. Meehl, C.D. Miller, S.J. Hassol, A.M. Waple, and W.L. Murray(eds.)]. Department of Commerce, NOAA s National Climatic Data Center, Washington, D.C., USA, pp. 164. Cervino, J.M., F.L. Thompson, B. Gomez-Gil, E.A. Lorence, T.J. Goreau, R.L. Hayes, K.B. Winiarski-Cervino, G.W. Smith, K. Hughen, and E. Bartels, 2008: The Vibrio core group induces yellow band disease in Caribbean and Indo-Pacific reef-building corals. Journal of Applied Microbiology, 105(5), 1658-1671. Chadee, D., B. Shivnauth, S. Rawlins, and A. Chen, 2007: Climate, mosquito indices and the epidemiology of dengue fever in Trinidad (2002 2004). Annals of Tropical Medicine and Parasitology, 101(1), 69-77. Chadee, D.D., 2009: Dengue cases and Aedes aegypti indices in Trinidad, West Indies. Acta Tropica, 112(2), 174- 180. Chan, W.H., Y.L. Mak, J.J. Wu, L. Jin, W.H. Sit, J.C.W. Lam, Y. Sadovy de Mitcheson, L.L. Chan, P.K.S. Lam, and M.B. Murphy, 2011: Spatial distribution of ciguateric fish in the Republic of Kiribati. Chemosphere, 84(1), 117-123. Charlery, J. and L. Nurse, 2010: Areal downscaling of global climate models: An approach that avoids data remodelling. Climate Research, 43(3), 2412-249. Chaves, L.F. and J.M. Koenraadt, 2010: Climate change and highland malaria: Fresh air in a hot debate . The Quarterly Review of Biology, 85(1), 27-55. Chen, F., N. Duic, L.M. Alves, and M.G. Carvalho, 2007: Renewislands Renewable energy solutions for islands. Renewable and Sustainable Energy Reviews, 8(11), 1888-1902. Chen, I., H.J. Shiu, S. Benedick, J.D. Holloway, V.K. Chey, H.S. Barlow, J.K. Hill, and C.D. Thomas, 2009: Elevation increases in moth assemblages over 42 years on a tropical mountain. Proceedings of the National Academy of Sciences, 106(5), 1479. Christenhusz, M.J.M. and T.K. Toivonen, 2008: Giants invading the tropics: the oriental vessel fern, Angiopteris evecta (Marattiaceae). Biological Invasions, 10(8), 1215-1228. Church, J.A. and N.J. White, 2011: Sea-Level Rise from the Late 19th to the Early 21st Century. Surveys in Geophysics, 32(4-5), 585-602. Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cinner, J.E., T.R. McClanahan, N.A.J. Graham, T.M. Daw, J. Maina, S.M. Stead, A. Wamukota, K. Brown, and Ö. Bodin, 2012: Vulnerability of coastal communities to key impacts of climate change on coral reef fisheries. Global Environmental Change, 22(1), 12-20. Cochard, R., S.L. Ranamukhaarachchi, G.P. Shivakoti, O.V. Shipin, P.J. Edwards, and K.T. Seeland, 2008: The 2004 tsunami in Aceh and Southern Thailand: A review on coastal ecosystems, wave hazards and vulnerability. Perspectives in Plant Ecology Evolution and Systematics, 10(1), 3-40. Connell, J. and J.P. Lea, 2002: Urbanisation in the Island Pacific: Towards Sustainable Development. Routledge, London, pp. 256. Connell, J., 2012: Population Resettlement in the Pacific: lessons from a hazardous history? Australian Geographer, 43(2), 127-142. Cooper, A., D. Jackson, and S. Gore, 2013: A groundswell event on the coast of the British Virgin Islands: variability in morphological impact. Journal of Coastal Research, SI65, 696-701. Coutinho, R., J. Pacheco, N. Wallenstein, A. Pimentel, R. Marques, and R. Silva, 2009: Integrating Geological Knowledge in Planning Methods for Small Islands Coastal Plans. Journal of Coastal Research, SI 56, 1199- 1203. Cronk, Q., 1997: Islands: stability, diversity, conservation. Biodiversity & Conservation, 6(3), 477-493. Crump, J., 2008: Snow, sand, ice and sun: Climate change and equity in the Arctic and small island developing states. Sustainable Development Law and Policy, 8(3), 8-13. Cruz, M., R. Aguiar, A. Correia, T. Tavares, J. Pereira, and F. Santos, 2009: Impacts of climate change on the terrestrial ecosystems of Madeira. International Journal of Design and Nature and Ecodynamics, 4(4), 413-422. Cummins, J.D. and O. Mahul, 2009: Catastrophe risk financing in developing countries: Principles for Public Intervention. World Bank Publications, Washington D.C., pp. 268. Danovaro, R., S.F. Umani, and A. Pusceddu, 2009: Climate Change and the potential spreading of marine mucilage and microbial pathogens in the Mediterranean Sea. Plos One, 4(9), e7006. Dawson, J.L. and S.G. Smithers, 2010: Shoreline and beach volume change between 1967 and 2007 at Raine Island, Great Barrier Reef, Australia. Global and Planetary Change, 72(3), 141-154. de Haas, H., 2011: Mediterranean migration futures: Patterns, drivers and scenarios. Global Environmental Change, 21, Supplement 1, S59-S69. de Scally, F.A., 2008: Historical tropical cyclone activity and impacts in the Cook Islands. Pacific Science, 62(4), 443-459. De'ath, G., J.M. Lough, and K.C. Fabricius, 2009: Declining coral calcification on the Great Barrier Reef. Science, 323(5910), 116-119. Dema, B., 2012: Sea Level Rise and the Freely Associated States: Addressing Environmental Migration Under the Compacts of Free Association. Columbia Journal of Environmental Law, 37, 177-204. Department of Meteorology (Maldives), 2007: The Unusually Strong Tidal Waves Hit Maldives, 15-18 May. Department of Meteorology, Republic of Maldives, pp. 12. Didham, R.K., J.M. Tylianakis, M.A. Hutchison, R.M. Ewers, and N.J. Gemmell, 2005: Are invasive species the drivers of ecological change? Trends in Ecology & Evolution, 20(9), 470-474. Donelson, J.M., P.L. Munday, M.I. McCormick, N.W. Pankhurst, and P.M. Pankhurst, 2010: Effects of elevated water temperature and food availability on the reproductive performance of a coral reef fish. Marine Ecology Progress Series, 401, 233-243. Donn, W.L. and W.T. McGuinness, 1959: Barbados storm swell. Journal of Geophysical Research, 64(12), 2341- 2349. Donner, S.D., 2009: Coping with commitment: Projected thermal stress on coral reefs under different Future scenarios. Plos One, 4(6), e5712. Donner, S.D., T. Kirata, and C. Vieux, 2010: Recovery from the 2004 coral bleaching event in the Gilbert Islands, Kiribati. Atoll Research Bulletin, (587), 1-25. Dornan, M., 2009: Methods for assessing the contribution of renewable technologies to energy security: the electricity sector of Fiji. Pacific Economic Bulletin, 24(3), 71-91. Dornan, M., 2011: Solar-based rural electrification policy design: The Renewable Energy Service Company (RESCO) model in Fiji. Renewable Energy, 36(2), 797-803. Dorville, J.F. and N. Zahibo, 2010: Hurricane Omar Waves Impact on the West Coast of the Guadeloupe Island, October 2008. The Open Oceanography Journal, 4, 83-91. Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Dulal, H.B., K.U. Shah, and N. Ahmad, 2009: Social Equity Considerations in the Implementation of Caribbean Climate Change Adaptation Policies. Sustainability, 1(3), 363-383. Dumaru, P., 2010: Community-based adaptation: enhancing community adaptive capacity in Druadrua Island, Fiji. Wiley Interdisciplinary Reviews Climate Change, 1(5), 751-763. Dunne, R.P., S.M. Barbosa, and P.L. Woodworth, 2012: Contemporary sea level in the Chagos Archipelago, central Indian Ocean. Global and Planetary Change, 82 83(0), 25-37. Dupon, J., 1987: Les atolls et le risque cyclonique. Cah.Sci.Hum, 23(5-4), 567-599. Duvat, V., 2009: Beach erosion management in small island developing states: Indian Ocean case studies. Proceedings of the International conference on physical coastal processes, management and engineering, September 14 - 16, Malta, pp. 149-160. Duvat, V., 2013: Coastal protection structures in Tarawa Atoll, Republic of Kiribati. Sustainability Science, 8(3), 363-379. Eakin, C.M., J.M. Lough, and S.F. Heron, 2009: Climate variability and change: Monitoring data and evidence for increased coral bleaching stress. Ecological Studies, 205, 41-67. Ellison, J.C., 1993: Mangrove Retreat with Rising Sea-Level, Bermuda. Estuarine, Coastal and Shelf Science, 37(1), 75-87. Ellison, J.C., 2009: Wetlands of the Pacific Island region. Wetlands Ecology and Management, 17(3), 169-206. Emmanuel, K. and B. Spence, 2009: Climate change implications for water resource management in Caribbean tourism. Worldwide Hospitality and Tourism Themes, 1(3), 252-268. Englund, R.A., 2008: Invasive species threats to native aquatic insect biodiversity and conservation measures in Hawai'i and French Polynesia. Journal of Insect Conservation, 12(3-4), 415-428. Etienne, S., 2012: Marine inundation hazards in French Polynesia: geomorphic impacts of Tropical Cyclone Oli in February 2010. Geological Society, London, Special Publications, 361(1), 21-39. Etienne, S. and J.P. Terry, 2012: Coral boulders, gravel tongues and sand sheets: Features of coastal accretion and sediment nourishment by Cyclone Tomas (March 2010) on Taveuni Island, Fiji. Geomorphology, 175 176(0), 54-65. Fazey, I., N. Pettorelli, J. Kenter, D. Wagatora, and D. Schuett, 2011: Maladaptive trajectories of change in Makira, Solomon Islands. Global Environmental Change, 21(4), 1275-1289. Feagin, R.A., N. Mukherjee, K. Shanker, A.H. Baird, J. Cinner, A.M. Kerr, N. Koedam, A. Sridhar, R. Arthur, L.P. Jayatissa, D. Lo Seen, M. Menon, S. Rodriguez, M. Shamsuddoha, and F. Dahdouh-Guebas, 2010: Shelter from the storm? Use and misuse of coastal vegetation bioshields for managing natural disasters. Conservation Letters, 3(1), 1-11. Ferdinand, I., G. O Brien, P. O Keefe, and J. Jayawickrama, 2012: The double bind of poverty and community disaster risk reduction: A case study from the Caribbean. International Journal of Disaster Risk Reduction, 2(0), 84-94. Figueroa, J., 2008: The need to strengthen malaria control in the Caribbean in the era of HIV/AIDS. West Indian Medical Journal, 57(5), 425-426. Fish, M.R., I.M. Cote, J.A. Horrocks, B. Mulligan, A.R. Watkinson, and A.P. Jones, 2008: Construction setback regulations and sea-level rise: Mitigating sea turtle nesting beach loss. Ocean & Coastal Management, 51(4), 330-341. Fletcher, C.H., C. Bochicchio, C. Conger, M. Engels, F. Feirstein, N. Frazer, C. Glenn, R.W. Grigg, E.E. Grossman, and J.N. Harvey, 2008: Geology of Hawaii reefs. In: Coral Reefs of the USA [Riegl, B. and R.E. Dodge(eds.)]. Springer, pp. 435-487. Forbes, D.L., T.S. James, M. Sutherland, and S.E. Nichols, 2013: Physical basis of coastal adaptation on tropical small islands. Sustainability Science, 8, 327-344. Ford, M., 2012: Shoreline Changes on an urban atoll in the Central Pacific Ocean: Majuro Atoll, Marshall Islands. Journal of Coastal Research, 28(1), 11-22. Ford, M., 2013: Shoreline changes interpreted from multi-temporal aerial photographs and high resolution satellite images: Wotje Atoll, Marshall Islands. Remote Sensing of Environment, 135, 130-140. Forster, J., P.W. Schuhmann, I.R. Lake, A.R. Watkinson, and J.A. Gill, 2012: The influence of hurricane risk on tourist destination choice in the Caribbean. Climatic Change, 114(3-4), 745-768. Fox, J.C., C.K. Yosi, P. Nimiago, F. Oavika, J.N. Pokana, K. Lavong, and R.J. Keenan, 2010: Assessment of above ground carbon in primary and selectively harvested tropical forest in Papua New Guinea. Biotropica, 42(4), 410-419. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Freed, L.A., R.L. Cann, M.L. Goff, W.A. Kuntz, and G.R. Bodner, 2005: Increase in avian malaria at upper elevations in Hawaii. The Condor, 107, 753-764. Frieler, K., Meinshausen M., A. Golly, M. Mengel, K. Lebek, Donner S. D., and Hoegh-Guldberg O., 2013: Limiting global warming to 2 °C is unlikely to save most coral reefs. Nature Climate Change, 3, 165-170. Gaigher, R., M. Samways, J. Henwood, and K. Jolliffe, 2011: Impact of a mutualism between an invasive ant and honeydew-producing insects on a functionally important tree on a tropical island. Biological Invasions, 13(8), 1717-1721. Gamble, D.W., D. Campbell, T.L. Allen, D. Barker, S. Curtis, D. McGregor, and J. Popke, 2010: Climate change, drought, and Jamaican agriculture: Local knowledge and the climate record. Annals of the Association of American Geographers, 100(4), 880-893. Game, E.T., G. Lipsett Moore, R. Hamilton, N. Peterson, J. Kereseka, W. Atu, M. Watts, and H. Possingham, 2011: Informed opportunism for conservation planning in the Solomon Islands. Conservation Letters, 4(1), 38-46. Garrison, V., W. Foreman, S. Genualdi, D. Griffin, C. Kellogg, M. Majewski, A. Mohammed, A. Ramsubhag, E. Shinn, and S. Simonich, 2006: Saharan dust-a carrier of persistent organic pollutants, metals and microbes to the Caribbean? Revista De Biología Tropical, 54, 9-21. Gemenne, F., 2011: Climate-induced population displacements in a 4 °C world. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1934), 182-195. Gero, A., K. Méheux, and D. Dominey-Howes, 2011: Integrating community based disaster risk reduction and climate change adaptation: examples from the Pacific. Natural Hazards and Earth System Sciences, 11(1), 101- 113. Gillespie, T.W., J. Chu, and S. Pau, 2008: Non-native plant invasion of the Hawaiian Islands. Geography Compass, 2(5), 1241-1265. Gilman, E.L., J. Ellison, N.C. Duke, and C. Field, 2008: Threats to mangroves from climate change and adaptation options: A review. Aquatic Botany, 89(2), 237-250. Gilmour, J.P., L.D. Smith, A.J. Heyward, A.H. Baird, and M.S. Pratchett, 2013: Recovery of an isolated coral reef system following severe disturbance. Science, 340(6128), 69-71. Goodman, J., J. Maschinski, P. Hughes, J. McAuliffe, J. Roncal, D. Powell, and L.O. Sternberg, 2012: Differential response to soil salinity in endangered key tree cactus: Implications for survival in a changing climate. Plos One, 7(3), e32528. Gordon-Clark, M., 2012: Paradise lost? Pacific island archives threatened by climate change. Archival Science, 12(1), 51-67. Gössling, S., O. Lindén, J. Helmersson, J. Liljenberg, and S. Quarm, 2012a: Diving and global environmental change: a Mauritius case study. In: New Frontiers in Marine Tourism: Diving Experiences, Management and Sustainability. [Garrod, B. and S. Gössling(eds.)]. Elsevier, Amsterdam, pp. 67. Gössling, S., P. Peeters, C.M. Hall, J. Ceron, G. Dubois, L.V. Lehmann, and D. Scott, 2012b: Tourism and water use: Supply, demand, and security. An international review. Tourism Management, 33(1), 1-15. Gössling, S. and K.P. Schumacher, 2010: Implementing carbon neutral destination policies: issues from the Seychelles. Journal of Sustainable Tourism, 18(3), 377-391. Gravelle, G. and N. Mimura, 2008: Vulnerability assessment of sea-level rise in Viti Levu, Fiji Islands. Sustainability Science, 3(2), 171-180. Greaver, T.L. and L.S.L. Sternberg, 2010: Decreased precipitation exacerbates the effects of sea level on coastal dune ecosystems in open ocean islands. Global Change Biology, 16(6), 1860-1869. Green, S.J., J.L. Akins, and I.M. Côté, 2011: Foraging behaviour and prey consumption in the Indo-Pacific lionfish on Bahamian coral reefs. Marine Ecology Progress Series, 433, 159-167. Green, S.J., J.L. Akins, A. Maljkoviæ, and I.M. Côté, 2012: Invasive lionfish drive Atlantic coral reef fish declines. PLoS One, 7(3), e32596. Griffin, D.W., 2007: Atmospheric movement of microorganisms in clouds of desert dust and implications for human health. Clinical Microbiology Reviews, 20(3), 459-477. Gritti, E., B. Smith, and M. Sykes, 2006: Vulnerability of Mediterranean Basin ecosystems to climate change and invasion by exotic plant species. Journal of Biogeography, 33(1), 145-157. Guillaumont, P., 2010: Assessing the economic vulnerability of small island developing states and the least developed countries. The Journal of Development Studies, 46(5), 828-854. Haigh, I.D., M. Eliot, and C. Pattiaratchi, 2011: Global influences of the 18.61 year nodal cycle and 8.85 year cycle of lunar perigee on high tidal levels. Journal of Geophysical Research, 116(C6), C06025. Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hansen, L., J. Hoffman, C. Drews, and E. Mielbrecht, 2010: Designing climate-smart conservation: Guidance and case Studies. Conservation Biology, 24(1), 63-69. Harangozo, S.A., 1992: Flooding in the Maldives and its implications for the global sea level rise debate. In: Sea Level Changes: Determination and Effects. [Woodworth, P.L., D.T. Pugh, J.G. DeRonde, R.G. Warrick, and J. Hannah(eds.)]. International Union of Geodesy and Geophysics and the American Geophysical Union, pp. 95- 99. Hardman, E.R., N.S. Stampfli, L. Hunt, S. Perrine, A. Perry, and J.S.J. Raffin, 2007: The impacts of coral bleaching in Rodrigues, Western Indian Ocean. Atoll Reseach Bulletin, (555), 10. Hare, W.L., W. Cramer, M. Schaeffer, A. Battaglini, and C.C. Jaeger, 2011: Climate hotspots: key vulnerable regions, climate change and limits to warming. Regional Environmental Change, 11(Supplement 1), S1-S13. Harmsen, E.W., N.L. Miller, N.J. Schlegel, and J.E. Gonzalez, 2009: Seasonal climate change impacts on evapotranspiration, precipitation deficit and crop yield in Puerto Rico. Agricultural Water Management, 96, 1085-1095. Harp, E.L., M.E. Reid, J.P. McKenna, and J.A. Michael, 2009: Mapping of hazard from rainfall-triggered landslides in developing countries: Examples from Honduras and Micronesia. Engineering Geology, 104(3-4), 295-311. Harris, A., G. Manahira, A. Sheppard, and Gough, C., Sheppard, C, 2010: Demise of Madagascar s once great barrier reef: changes in coral reef conditions over 40 years. Atoll Reseach Bulletin, (574), 1-16. Hay, J.E., 2013: Small island developing states: coastal systems, global change and sustainability. Sustainability Science, 8, 309-326. Hay, J.E. and N. Mimura, 2013: Vulnerability, risk and adaptation assessment methods in the Pacific Islands Region: past approaches, and considerations for the future. Sustainability Science, 8, 391-405. Hay, S.I., C.A. Guerra, P.W. Gething, A.P. Patil, A.J. Tatem, A.M. Noor, C.W. Kabaria, B.H. Manh, I.R. Elyazar, and S. Brooker, 2009: A world malaria map: Plasmodium falciparum endemicity in 2007. PLoS Medicine, 6(3), e1000048. Heger, M., A. Julca, and O. Paddison, 2008: Analysing the impact of natural hazards in small economies: The Caribbean case. UNU-WIDER research paper, No. 2008/25. Heltberg, R., P.B. Siegel, and S.L. Jorgensen, 2009: Addressing human vulnerability to climate change: Toward a 'no-regrets' approach. Global Environmental Change, 19(1), 89-99. Hemer, M.A., Y. Fan, N. Mori, A. Semedo, and X.L. Wang, 2013: Projected changes in wave climate from a multi- model ensemble. Nature Climate Change, 3(5), 471-476. Herrmann-Storck, C., D. Postic, I. Lamaury, and J.M. Perez, 2008: Changes in epidemiology of leptospirosis in 2003-2004, a two El nino Southern Oscillation period, Guadeloupe archipelago, French West Indies. Epidemiology and Infection, 136(10), 1407-1415. Herweijer, C., N. Ranger, and R.E.T. Ward, 2009: Adaptation to climate change: threats and opportunities for the insurance industry. The Geneva Papers on Risk and Insurance-Issues and Practice, 34(3), 360-380. Hoegh-Guldberg, O., P.J. Mumby, A.J. Hooten, R.S. Steneck, P. Greenfield, E. Gomez, C.D. Harvell, P.F.H. Sale, A. Dubi, and M.E. Hatziolos, 2007: Coral reefs under rapid climate change and ocean acidification. Science, 318(5857), 1737-1742. Hoeke, R.K., K.L. McInnes, J. Kruger, R. McNaught, J. Hunter, and S. Smithers, 2013: Widespread inundation of Pacific islands by distant-source wind-waves. Global Environmental Change, 108, 128-138. Hornidge, A. and F. Scholtes, 2011: Climate Change and Everyday Life in Toineke Village, West Timor: Uncertainties, Knowledge and Adaptation. Sociologus, 61(2), 151-175. Houston, J.R., 2002: The economic value of beaches-a 2002 update. Shore and Beach, 70(1), 9-12 Howard, G., K. Charles, K. Pond, A. Brookshaw, R. Hossain, and J. Bartram, 2010: Securing 2020 vision for 2030: climate change and ensuring resilience in water and sanitation services. Journal of Water and Climate Change, 1(1), 2-16. Hughes, S., A. Yau, L. Max, N. Petrovic, F. Davenport, M. Marshall, T.R. McClanahan, E.H. Allison, and J.E. Cinner, 2012: A framework to assess national level vulnerability from the perspective of food security: The case of coral reef fisheries. Environmental Science & Policy, 23, 95-108. Hughes, T.P., A.H. Baird, D.R. Bellwood, M. Card, S.R. Connolly, C. Folke, R. Grosberg, O. Hoegh-Guldberg, J. Jackson, J. Kleypas, J.M. Lough, P. Marshall, M. Nyström, S.R. S. R. Palumbi, J.M. J. M. Pandolfi, B. Rosen, and J. Roughgarden, 2003: Climate change, human impacts, and the resilience of coral reefs. Science, 301(5635), 929-933. Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hübner, A. and S. Gössling, 2012: Tourist perceptions of extreme weather events in Martinique. Journal of Destination Marketing & Management, 1, 47-55. Idjadi, J.A., R.N. Haring, and W.F. Precht, 2010: Recovery of the sea urchin Diadema antillarum promotes scleractinian coral growth and survivorship on shallow Jamaican reefs. Marine Ecology-Progress Series, 403, 91-100. IPCC (ed.), 2011: IPCC Special Report on Renewable Energy Sources and Climate Change Mitigation. Prepared by Working Group III of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, pp. 1075. Jarvis, R., 2010: Sinking nations and climate change adaptation strategies. Seattle Journal of Social Justice, 9(1), 447-486. Joe, S.M. and C.C. Daehler, 2008: Invasive slugs as under-appreciated obstacles to rare plant restoration: evidence from the Hawaiian Islands. Biological Invasions, 10(2), 245-255. Jones, G.P., M.I. McCormick, M. Srinivasan, and J.V. Eagle, 2004: Coral decline threatens fish biodiversity in marine reserves. Proceedings of the National Academy of Sciences, 101(21), 8251-8253. Jones, P., 2005: Managing Urban development in the Pacific. Australian Planner, 42(1), 39-46. Jury, M.R. and A. Winter, 2010: Warming of an elevated layer over the Caribbean. Climatic Change, 99, 247-259. Kaldellis, J.K., 2008: Integrated electrification solution for autonomous electrical networks on the basis of RES and energy storage configurations. Energy Conversion and Management, 49(12), 3708-3720. Kaldellis, J.K., D. Zafirakis, E.L. Kaldelli, and K. Kavadias, 2009: Cost benefit analysis of a photovoltaic-energy storage electrification solution for remote islands. Renewable Energy, 34(5), 1299-1311. Keener, V., J.J. Marra, M.L. Finucane, D. Spooner, and M.H. Smith, 2012: Climate Change and Pacific Islands: Indicators and Impacts: Report for the 2012 Pacific Islands Regional Climate Assessment. Island Press, Washington, DC, pp. 170. Kelman, I. and J. West, 2009: Climate change and small island developing states: A critical review. Ecological and Environmental Anthropology, 5(1), 1-16. Kelman, I., J. Lewis, J.C. Gaillard, and J. Mercer, 2011: Participatory action research for dealing with disasters on islands. Island Studies Journal, 6(1), 59-86. Kelman, I., 2010: Introduction to climate, disasters and international development. Journal of International Development, 22(2), 208-217. Kenis, M., M.A. Auger-Rozenberg, A. Roques, L. Timms, C. Pere, M. Cock, J. Settele, S. Augustin, and C. Lopez- Vaamonde, 2009: Ecological effects of invasive alien insects. Biological Invasions, 11(1), 21-45. Kingsford, R., J. Watson, C. Lundquist, O. Venter, L. Hughes, E. Johnston, J. Atherton, M. Gawel, D. Keith, B. Mackey, C. Morley, H.B. Possingham, B. Raynor, H.F. Recher, and K.A. Wilson, 2009: Major conservation policy issues for biodiversity in Oceania. Conservation Biology, 23(4), 834-840. Klein, R., S. Ericksen, L. Ness, A. Hammill, T. Tanner, C. Robledo, and K. O'Brien, 2007: Portfolio screening to support the mainstreaming of adaptation to climate change into development assistance. Climatic Change, 84(1), 23-44. Klint, L.M., M. Jiang, A. Law, T. Delacy, S. Filep, E. Calgaro, D. Dominey-Howes, and D. Harrison, 2012: Dive Tourism in Luganville, Vanuatu: Shocks, Stressors, and Vulnerability to Climate Change. Tourism in Marine Environments, 8(1-2), 91-109. Koh, B.K., L.C. Ng, Y. Kita, C.S. Tang, L.W. Ang, K.Y. Wong, L. James, and K.T. Goh, 2008: The 2005 dengue epidemic in Singapore: epidemiology, prevention and control. Annals Academy of Medicine Singapore, 37(7), 538. Kolmannskog, V. and L. Trebbi, 2010: Climate change, natural disasters and displacement: a multi-track approach to filling the protection gaps International Review of the Red Cross, 92(879), 713-730. Komar, P.D. and J.C. Allan, 2008: Increasing hurricane-generated wave heights along the US East coast and their climate controls. Journal of Coastal Research, 24(2), 479-488. Krajaèiæ, G., N. Duiæ, B. Vad Mathiesen, and M.D.G. Carvalho, 2010: Smart energy storages for integration of renewables in 100% independent energy systems. Chemical Engineering Transactions, 21, 391-396. Krauss, K.W., J.A. Allen, and D.R. Cahoon, 2003: Differential rates of vertical accretion and elevation change among aerial root types in Micronesian mangrove forests. Estuarine, Coastal and Shelf Science., 56, 251-259. Krauss, K.W., D.R. Cahoon, J.A. Allen, K.C. Ewel, J.C. Lynch, and N. Cormier, 2010: Surface elevation change and susceptibility of different mangrove zones to sea-level rise on Pacific high islands of Micronesia. Ecosystems, 13(1), 129-143. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kravchenko, S., 2008: Right to Carbon or Right to Life: Human Rights Approaches to Climate Change. Vermont Journal of Environmental Law, 9(3), 513-547. Krishnamurthy, P.K., 2012: Disaster-induced migration: Assessing the impact of extreme weather events on livelihoods. Environmental Hazards, 11(2), 96-111. Krushelnycky, P.D., L.L. Loope, T.W. Giambelluca, F. Starr, K. Starr, D.R. Drake, A.D. Taylor, and R.H. Robichaux, 2013: Climate associated population declines reverse recovery and threaten future of an iconic high elevation plant. Global Change Biology, 19(3), 911-922. Kueffer, C., C.C. Daehler, C.W. Torres-Santana, C. Lavergne, J.Y. Meyer, R. Otto, and L. Silva, 2010: A global comparison of plant invasions on oceanic islands. Perspectives in Plant Ecology Evolution and Systematics, 12(2), 145-161. Kumar, V.S., V.R. Babu, M. Babu, G. Dhinakaran, and G. Rajamanickam, 2008: Assessment of storm surge disaster potential for the Andaman Islands. Journal of Coastal Research, 24(2A), 171-177. Kuruppu, N., 2009: Adapting water resources to climate change in Kiribati: the importance of cultural values and meanings. Environmental Science & Policy, 12(7), 799-809. Kuruppu, N. and D. Liverman, 2011: Mental preparation for climate adaptation: The role of cognition and culture in enhancing adaptive capacity of water management in Kiribati. Global Environmental Change, 21(2), 657-669. Laczko, F. and C. Aghazarm, 2009: Migration, environment and climate change: assessing the evidence. International Organization for Migration, Geneva, Switzerland, . Lata, S. and P. Nunn, 2012: Misperceptions of climate-change risk as barriers to climate-change adaptation: a case study from the Rewa Delta, Fiji. Climatic Change, 110(1-2), 169-186. Lauer, M. and S. Aswani, 2010: Indigenous knowledge and long-term ecological change: detection, interpretation, and responses to changing ecological conditions in pacific Island Communities. Environmental Management, 45(5), 985-997. Lazrus, H., 2012: Sea Change: Island Communities and Climate Change. Annual Review of Anthropology, 41(1), 285-301. Le Masson, V. and I. Kelman, 2011: Disaster risk reduction on non-sovereign islands: La Réunion and Mayotte, France. Natural Hazards, 56(1), 251-273. Le Roux, P.C., M.A. McGeoch, M.J. Nyakatya, and S.L. Chown, 2005: Effects of a short term climate change experiment on a sub Antarctic keystone plant species. Global Change Biology, 11(10), 1628-1639. Le Roux, J.J., A.M. Wieczorek, and J.Y. Meyer, 2008: Genetic diversity and structure of the invasive tree Miconia calvescens in Pacific islands. Diversity and Distributions, 14(6), 935-948. Lefale, P., 2010: Ua afa le Aso stormy weather today: traditional ecological knowledge of weather and climate. The Samoa experience. Climatic Change, 100(2), 317-335. Lesser, M.P. and M. Slattery, 2011: Phase shift to algal dominated communities at mesophotic depths associated with lionfish (Pterois volitans) invasion on a Bahamian coral reef. Biological Invasions, 13(8), 1855-1868. Lessios, H.A., 1988: Mass mortality of Diadema antillarum in the Caribbean: What have we learned? Annual Review of Ecology and Systematics, 19(1), 371-393. Lessios, H.A., 1995: Diadema antillarum 10 years after mass mortality: Still rare, despite help from a competitor. Proceedings of the Royal Society of London. Series B: Biological Sciences, 259(1356), 331-337. Liller, H.B. and K. Van den Broeck, 2011: Economic drivers of migration and climate change in LDCs. Global Environmental Change, 21, Supplement 1(0), S70-S81. Limalevu, L., B. Aalsbersberg, P. Dumaru, and T. Weir, 2010: Adaptation on a small island. Tiempo, 77, 16. Linnerooth-Bayer, J., R. Mechler, and S. Hochrainer, 2011: Insurance against Losses from Natural Disasters in Developing Countries. Evidence, gaps and the way forward. Journal of Integrated Disaster Risk Management, 1(1), 1-23. Linnerooth-Bayer, J. and R. Mechler, 2009: Insurance against Losses from Natural Disasters in Developing Countries. DESA Working Paper No. 85. UN/DESA Working Papers, New York, USA, pp. 1-35. Llewellyn, L.E., 2010: Revisiting the association between sea surface temperature and the epidemiology of fish poisoning in the South Pacific: Reassessing the link between ciguatera and climate change. Toxicon, 56(5), 691- 697. Locke, J.T., 2009: Climate change-induced migration in the Pacific Region: sudden crisis and long-term developments. The Geographical Journal, 175(3), 171-180. López-Marrero, T., 2010: An integrative approach to study and promote natural hazards adaptive capacity: a case study of two flood-prone communities in Puerto Rico. Geographical Journal, 176(2), 150-163. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 López-Marrero, T. and B. Yarnal, 2010: Putting adaptive capacity into the context of people s lives: a case study of two flood-prone communities in Puerto Rico. Natural Hazards, 52(2), 277-297. López-Villarrubia, E., F. Ballester, C. Iniguez, and N. Peral, 2010: Air pollution and mortality in the Canary Islands: a time-series analysis. Environ Health, 9(8), 1-11. Loughry, M. and J. McAdam, 2008: Kiribati relocation and adaptation. Forced Migration Review, 31, 51-52. Lovell, S., 2011: Health governance and the impact of climate change on Pacific small island developing states. In: IHDP Update (Human Health and Global Environmental Change). International Human Dimensions Program, pp. 50-55. Lozano, M.A., 2006: Riesgos Catastróficos en las Islas Canarias. Una Visión Geográfica. Anales De Geografía, 26, 167-194. Lo-Yat, A., S.D. Simpson, M. Meekan, D. Lecchini, E. Martinez, and R. Galzin, 2011: Extreme climatic events reduce ocean productivity and larval supply in a tropical reef ecosystem. Global Change Biology, 17(4), 1695- 1702. Maina, J., T.R. McClanahan, V. Venus, M. Ateweberhan, and J. Madin, 2011: Global gradients of coral exposure to environmental stresses and implications for local management. PLoS One, 6(8), e23064. Maniku, H.A., 1990: Changes in the Topography of the Maldives. Forum of Writers on Environment (Maldives), Male, pp. 91. Marba, N. and C.M. Duarte, 2010: Mediterranean warming triggers seagrass (Posidonia oceanica) shoot mortality. Global Change Biology, 16(8), 2366-2375. Martens, P., R.S. Kovats, S. Nijof, P. deVries, M.T.J. Livermore, D.J. Bradley, and et al., Global environmental Change: Ciate change and future populations at risk of malaria. 1999, 9(Suppl 1), S89-S107. Martins, P., J. Rosado-Pinto, M. Do Céu Teixeira, N. Neuparth, O. Silva, H. Tavares, J.L. Spencer, D. Mascarenhas, A.L. Papoila, N. Khaltaev, and I. Annesi-Maesano, 2009: Under-report and underdiagnosis of chronic respiratory diseases in an African country. Allergy, 64(7), 1061-1067. Maunsell Australia Pty Ltd., 2009: Climate Change Risk Assessment for the Australian Indian Ocean Territories. Commonwealth Attorney General's Department, Canberra, Australia, pp. 1-93. McClanahan, T., J. Cinner, J. Maina, N. Graham, T. Daw, S. Stead, A. Wamukota, K. Brown, M. Ateweberhan, and V. Venus, 2008: Conservation action in a changing climate. Conservation Letters, 1(2), 53-59. McGranahan, G., D. Balk, and B. Anderson, 2007: The rising tide: assessing the risks of climate change and human settlements in low elevation coastal zones. Environment and Urbanization, 19(1), 17-37. McGray, H., A. Hammil, and R. Bradley, 2007: Weather the storm. Options for framing adaptation and development. World Resources Institute, Washington DC, pp. 57. McKee, K.L., D.R. Cahoon, and I.C. Feller, 2007: Caribbean mangroves adjust to rising sea level through biotic controls on change in soil elevation. Global Ecology and Biogeography, 16(5), 556. McLeman, R.A. and L.M. Hunter, 2010: Migration in the context of vulnerability and adaptation to climate change: Insights from analogues. Wiley Interdisciplinary Reviews Climate Change, 1(3), 450-461. McLeman, R.A., 2011: Settlement abandonment in the context of global environmental change. Global Environmental Change, 21, Supplement 1(0), S108-S120. McLeod, E., R. Salm, A. Green, and J. Almany, 2009: Designing marine protected area networks to address the impacts of climate change. Frontiers in Ecology and the Environment, 7(7), 362-370. McMichael, C., J. Barnett, and A.J. McMichael, 2012: An Ill Wind? Climate Change, Migration, and Health. Environmental Health Perspectives, 120(5), 646-654. McMichael, A.J. and E. Lindgren, 2011: Climate change: present and future risks to health, and necessary responses. Journal of Internal Medicine, 270(5), 401-413. McNamara, K.E. and C. Gibson, 2009: We do not want to leave our land : Pacific ambassadors at the United Nations resist the category of climate refugees . Geoforum, 40(3), 475-483. McNamara, K.E. and R. Westoby, 2011: Local knowledge and climate change adaptation on Erub Island, Torres Strait. Local Environment, 16(9), 887-901. Mercer, J., D. Dominey-Howes, I. Kelman, and K. Lloyd, 2007: The potential for combining indigenous and western knowledge in reducing vulnerability to environmental hazards in small island developing states. Environmental Hazards, 7(4), 245-256. Merrifield, M., S. Merrifield, and G. Mitchum, 2009: An anomalous recent acceleration of global sea level rise. Journal of Climate, 22(21), 5772-5781. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mertz, O., T.B. Bruun, B. Fog, K. Rasmussen, and J. Agergaard, 2010: Sustainable land use in Tikopia: Food production and consumption in an isolated agricultural system. Singapore Journal of Tropical Geography, 31(1), 10-26. Mertz, O., K. Halsnaes, J. Olesen, and K. Rasmussen, 2009: Adaptation to climate change in developing countries. Environmental Management, 43(5), 743-752. Metz, B., O.R. Davidson, P.R. Bosch, R. Dave, and L.A. Meyers (eds.), 2007: Mitigation. Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK; New York, NY, USA, . Meyer, W.M., K.A. Hayes, and A.L. Meyer, 2008: Giant African snail, Achatina fulica, as a snail predator. American Malacological Bulletin, 24(1-2), 117-119. Meyssignac, B., M. Becker, W. Llovel, and A. Cazenave, 2012: An assessment of two-dimensional past sea level recontructions over 1950-2009 based on tide-gauge data and different input sea level grids. Surveys in Geophysics., 33(945), 972. Michon, P., J.L. Cole-Tobian, E. Dabod, S. Schoepflin, J. Igu, M. Susapu, N. Tarongka, P.A. Zimmerman, J.C. Reeder, and J.G. Beeson, 2007: The risk of malarial infections and disease in Papua New Guinean children. The American Journal of Tropical Medicine and Hygiene, 76(6), 997-1008. Middleton, N., P. Yiallouros, S. Kleanthous, O. Kolokotroni, J. Schwartz, D.W. Dockery, P. Demokritou, and P. Koutrakis, 2008: A 10-year time-series analysis of respiratory and cardiovascular morbidity in Nicosia, Cyprus: The effect of short-term changes in air pollution and dust storms. Environmental Health, 7(39). Miller, J., E. Muller, C. Rogers, R. Waara, A. Atkinson, K.R.T. Whelan, M. Patterson, and B. Witcher, 2009: Coral disease following massive bleaching in 2005 causes 60% decline in coral cover on reefs in the US Virgin Islands. Coral Reefs, 28(4), 925-937. Miller, S., T. Brewer, and N. Harris, 2009: Rainfall thresholding and susceptibility assessment of rainfall-induced landslides: application to landslide management in St Thomas, Jamaica. Bulletin of Engineering Geology and the Environment, 68(4), 539-550. Mills, M., R.L. Pressey, R. Weeks, S. Foale, and N.C. Ban, 2010: A mismatch of scales: challenges in planning for implementation of marine protected areas in the Coral Triangle. Conservation Letters, 3(5), 291-303. Mimura, N., L.A. Nurse, R.F. McLean, J. Agard, P. Brigguglio, R. Lefale, R. Payet, and G. Sem, 2007: Small islands. In: Climate Change 2007: Impacts, Adaptation and Vulnerability Contribution of Working Group II in the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. [Parry, M., O.F. Canziani, J.P. Palutikof, P.J. van der Linden, and C.E. Hanson (eds.)]. Cambridge University Press, Cambridge, U.K., pp. 687-716. Moglia, M., P. Perez, and S. Burn, 2008a: Urbanization and water development in the Pacific Islands. Development, 51(1), 49-55. Moglia, M., P. Perez, and S. Burn, 2008b: Water troubles in a Pacific atoll town. Water Policy, 10(6), 613-637. Mohan, A., A. Cumberbatch, A. Adesiyun, and D.D. Chadee, 2009: Epidemiology of human leptospirosis in Trinidad and Tobago,1996-200: A retrospective study. Acta Tropica, 112(3), 260-265. Mohanty, M., 2012: New renewable energy sources, green energy development and climate change: Implications to Pacific Island countries. Management of Environmental Quality: An International Journal, 23(3), 264-274. Monteil, M., 2008: Saharan dust clouds and human health in the English-speaking Caribbean: what we know and don t know. Environmental Geochemistry and Health, 30(4), 339-343. Monteil, M. and R. Antoine, 2009: African dust and asthma in the Caribbean: medical and statistical perspectives. International Journal of Biometeorology, 53(5), 379-381. Moore, W.R., 2010: The impact of climate change on Caribbean tourism demand. Current Issues in Tourism, 13(5), 495-505. Moreno, A. and Becken.S., 2009: A climate change vulnerability assessment methodology for coastal tourism. Journal of Sustainable Tourism, 17(4), 473-488. Morrison, K., P. Prieto, A. Dominguez, D. Waltner-Toews, and J. FitzGibbon, 2008: Ciguatera Fish Poisoning in La Habana, Cuba: A Study of Local Social-Ecological Resilience. Ecohealth, 5(3), 346-359. Mortreux, C. and J. Barnett, 2009: Climate change, migration and adaptation in Funafuti, Tuvalu. Global Environmental Change, 19(1), 105-112. Moss, R.H., J.A. Edmonds, K.A.M. Hibbard M.R., S.K. Rose, D.P. van Vuuren, T.R. Carter, S. Emori, M. Kainuma, T. Kram, G.A. Meehl, J.F.B. Mitchell, N. Nakicenovic, K. Riahi, S.J. Smith, R.J. Stouffer, A.M. Thomson, J.P. Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Weyant, and T.J. Wilbanks, 2010: The next generation of scenarios for climate change research and assessment. Nature, 463, 747-756. Munday, P.L., J.M. Leis, J.M. Lough, C.B. Paris, M.J. Kingsford, M.L. Berumen, and J. Lambrechts, 2009: Climate change and coral reef connectivity. Coral Reefs, 28(2), 379-395. Mycoo, M., 2011: Natural hazard risk reduction: Making St. Lucia safe in an era of increased hurricanes and associated events. Natural Hazards Review, 12(1), 37-45. Mycoo, M., 2007: Diagnosis of Trinidad's water problems (mid-1980s to mid-1990s). Water Policy, 9(1), 73-86. Mycoo, M. and A. Chadwick, 2012: Adaptation to climate change: the coastal zone of Barbados. Proceedings of the ICE-Maritime Engineering, 165(4), 159-168. Myers, N., 2002: Environmental refugees: a growing phenomenon of the 21st century. Philosophical Transactions of the Royal Society of London.Series B: Biological Sciences, 357(1420), 609-613. Nelson, D.R., 2011: Adaptation and resilience: responding to a changing climate. Wiley Interdisciplinary Reviews: Climate Change, 2(1), 113-120. Nicholls, R.J. and A. Cazenave, 2010: Sea-level rise and its impact on coastal zones. Science, 328(5985), 1517- 1520. Nicholls, R.J. and R.S. Tol, 2006: Impacts and responses to sea-level rise: a global analysis of the SRES scenarios over the twenty-first century. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 364(1841), 1073-1095. Nicholls, R.J., N. Marinova, J.A. Lowe, S. Brown, P. Vellinga, D. De Gusma o, J. Hinkel, and R.S.J. Tol, 2011: Sea- level rise and its possible impacts given a 'beyond 4°C world' in the twenty-first century. Philosophical Transactions of the Royal Society A: Mathematical Physical and Engineering Sciences, 369(1934), 161-181. Nicoll, K.A., R.G. Harrison, and Z. Ulanowski, 2011: Observation of Saharan dust electrification. Environmental Research Letters, 6(1), 1-6. Novelo-Casanova, D.A. and Suarez G., 2010: Natural and man-made hazards in the Cayman Islands. Natural Hazards, 55(2), 441-466. Nunn, P.D., W. Aalbersberg, S. Lata, and M. Gwilliam, 2013: Beyond the core: community governance for climate- change adaptation in peripheral parts of Pacific Island Countries. Regional Environmental Change, DOI 10.1007/s10113-013-0486-7 (http://link.springer.com/article/10.1007%2Fs10113-013-0486-7#page-1). Nunn, P.D., 2009: Responding to the challenges of climate change in the Pacific Islands: management and technological imperatives. Climate Research, 40(2-3), 211-231. Nunn, P.D., 2007: Climate, Environment and Society in the Pacific During the Last Millennium. In: Developments in Earth and Environmental Sciences. Elsevier, pp. v-302. Nurse, L.A., G. Sem, J.E. Hay, A.G. Suarez, P.P. Wong, S. Briguglio, and S. Ragoonaden, 2001: Small island states. In: Climate Change 2001: Impacts, Adaptation and Vulnerability, Contribution of Working Group II to the Third Assessment of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, pp. 843-912. OECS (Organisation of East Caribbean States), 2004: Grenada: Macro Socio-economic Assessment of the Damages Caused by Hurricane Ivan. OECS, Castries, St. Lucia, pp. 139. Ogston, A.S. and M.E. Field, 2010: Predictions of turbidity due to enhanced sediment resuspension resulting from sea-level rise on a fringing coral reef: Evidence from Molokai, Hawaii. Journal of Coastal Research, 26(6), 1027-1037. Otero, P., S. Perez, A. Alfonso, C. Vale, P. Rodriguez, N.N. Gouveia, N. Gouveia, J. Delgado, P. Vale, M. Hirama, Y. Ishihara, J. Molgo, and L.M. Botana, 2010: First toxin profile of Ciguateric fish in Madeira Arquipelago (Europe). Analytical Chemistry, 82(14), 6032-6039. Oxenford, H.A., R. Roach, A. Brathwaite, L. Nurse, R. Goodridge, F. Hinds, K. Baldwin, and C. Finney, 2008: Quantitative observations of a major coral bleaching event in Barbados, Southeastern Caribbean. Climatic Change, 87(3-4), 435-449. Palanisamy, H., M. Becker, B. Meyssignac, O. Henry, and A. Cazenave, 2012: Regional sea level change and variability in the Caribbean sea since 1950. Journal of Geodetic Science, 2(2), 125-133. Pappas, G., P. Papadimitriou, V. Siozopoulou, L. Christou, and N. Akritidis, 2008: The globalization of leptospirosis: worldwide incidence trends. International Journal of Infectious Diseases, 12(4), 351-357. Parham, P.E. and E. Michael, 2010: Modelling the effects of weather and climate change on malaria transmission. Environmental Health Perspectives, 118(5), 620-626. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Park, S., N. Marshall, E. Jakku, A. Dowd, S. Howden, E. Mendham, and A. Fleming, 2012: Informing adaptation responses to climate change through theories of transformation. Global Environmental Change, 22(1), 115-126. Parks, B.C. and J.T. Roberts, 2006: Globalization, vulnerability to climate change, and perceived injustice. Society and Natural Resources, 19(4), 337-355. Parolo, G. and G. Rossi, 2008: Upward migration of vascular plants following a climate warming trend in the Alps. Basic and Applied Ecology, 9(2), 100-107. Pascual, M., J.A. Ahumada, L.F. Chaves, X. Rodo, and M. Bouma, 2006: Malaria resurgence in the East African highlands: temperature trends revisited. Proceedings of the National Academy of Sciences, 103(15), 5829-5834. Pauli, H., M. Gottfried, K. Reiter, C. Klettner, and G. Grabherr, 2007: Signals of range expansions and contractions of vascular plants in the high Alps: observations (1994 2004) at the GLORIA* master site Schrankogel, Tyrol, Austria. Global Change Biology, 13(1), 147-156. Payet, R. and W. Agricole, 2006: Climate change in the Seychelles: Implications for water and coral reefs. Ambio, 35(4), 182-189. Peduzzi, P., H. Dao, C. Herold, and F. Mouton, 2009: Assessing global exposure and vulnerability towards natural hazards: the Disaster Risk Index. Natural Hazards and Earth System Sciences, 9(4), 1149-1159. Pelling, M. and J.I. Uitto, 2001: Small island developing states: natural disaster vulnerability and global change. Global Environmental Change Part B: Environmental Hazards, 3(2), 49-62. Pelling, M., 2011: Urban governance and disaster risk reduction in the Caribbean: the experiences of Oxfam GB. Environment and Urbanization, 23(2), 383-400. Perch-Nielsen, S.L., 2010: The vulnerability of beach tourism to climate change an index approach. Climatic Change, 100(3-4), 579-606. Pérez-Arellano, J., O.P. Luzardo, A.P. Brito, M.H. Cabrera, M. Zumbado, C. Carranza, A. Angel-Moreno, R.W. Dickey, and L.D. Boada, 2005: Ciguatera Fish Poisoning, Canary Islands. Emerging Infectious Diseases, 11(12), 1981-1982. Peters, G.P., R.M. Andrew, T. Boden, J.G. Canadell, P. Ciais, C. Le Quéré, G. Marland, M.R. Raupach, and C. Wilson, 2013: The challenge to keep global warming below 2 °C. Nature Climate Change, 3, 4-6. Piskorska, M., G. Smith, and E. Weil, 2007: Bacteria associated with the coral Echinopora lamellosa (Esper 1795) in the Indian Ocean - Zanzibar Region. African Journal of Environmental Science and Technology, 1(5), 93-98. Polidoro, B.A., K.E. Carpenter, L. Collins, N.C. Duke, A.M. Ellison, J.C. Ellison, E.J. Farnsworth, E.S. Fernando, K. Kathiresan, N.E. Koedam, S.R. Livingstone, T. Miyagi, G.E. Moore, V.N. Nam, J.E. Ong, J.H. Primavera, S.G. Salmo, J.C. Sanciangco, S. Sukardjo, Y. Wang, and J.W.H. Yong, 2010: The Loss of Species: Mangrove Extinction Risk and Geographic Areas of Global Concern. PLoS One, 5(4), e10095. Polovina, J.J., J.P. Dunne, P.A. Woodworth, and E.A. Howell, 2011: Projected expansion of the subtropical biome and contraction of the temperate and equatorial upwelling biomes in the North Pacific under global warming. ICES Journal of Marine Science: Journal Du Conseil, 68(6), 986-995. Praene, J.P., M. David, F. Sinama, D. Morau, and O. Marc, 2012: Renewable energy: Progressing towards a net zero energy island, the case of Reunion Island. Renewable and Sustainable Energy Reviews, 16(1), 426-442. Prasad, B.C., 2008: Institutions, good governance and economic growth in the Pacific Island countries. International Journal of Social Economics, 35(12), 904-918. Pratchett, M.S., S.K. Wilson, A.J. Graham, P.L. Munday, G.P. Jones, and N.V.C. Polunin, 2009: Coral bleaching and consequences for motile reef organisms: past, present and uncertain future effects. Ecological Studies, 205, 139-158. Prevatt, D., L. Dupigny-Giroux, and F. Masters, 2010: Engineering Perspectives on Reducing Hurricane Damage to Housing in CARICOM Caribbean Islands. Natural Hazards Review, 11(4), 140-150. Prospero, J.M., 2006: Case study: Saharan dust impacts and climate change. Oceanography, 19(2), 59-61. Prospero, J., E. Blades, R. Naidu, G. Mathison, H. Thani, and M. Lavoie, 2008: Relationship between African dust carried in the Atlantic trade winds and surges in pediatric asthma attendances in the Caribbean. International Journal of Biometeorology, 52(8), 823-832. Prospero, J.M. and P.J. Lamb, 2003: African droughts and dust transport to the Caribbean: Climate change implications. Science, 302(5647), 1024-1027. Prospero, J.M., E. Blades, G. Mathison, and R. Naidu, 2005: Interhemispheric transport of viable fungi and bacteria from Africa to the Caribbean with soil dust. Aerobiologia, 21(1), 1-19. Pulwarty, R.S., L.A. Nurse, and U.O. Trotz, 2010: Caribbean Islands in a changing climate. Environment: Science and Policy for Sustainable Development, 52(6), 16-27. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ralph, P.J., M.J. Durako, S. Enriquez, C.J. Collier, and M.A. Doblin, 2007: Impact of light limitation on seagrasses. Journal of Experimental Marine Biology and Ecology, 350(1-2), 176-193. Rankey, E.C., 2011: Nature and stability of atoll island shorelines: Gilbert Island chain, Kiribati, equatorial Pacific. Sedimentology, 58(7), 1831-1859. Rasmussen, K., W. May, T. Birk, M. Mataki, and O. Mertz, 2011: Prospects for climate change on three Polynesian outliers in Solomon Islands: Exposure, sensitivity and adaptive capacity. Geografisk Tidsskrift-Danish Journal of Geography, 111(1), 43-57. Rasmussen, K., W. May, T. Birk, M. Mataki, O. Mertz, and D. Yee, 2009: Climate change on three Polynesian outliers in the Solomon Islands: Impacts, vulnerability and adaptation. Geografisk Tidsskrift Danish Journal of Geography, 109(1), 1-13. Rawlins, S.C., A. Hinds, and J.M. Rawlins, 2008: Malaria and its vectors in the Caribbean: The continuing challenge of the disease forty-five years after eradication from the islands. West Indian Medical Journal, 57(5), 462-469. Reaser, J.K., L.A. Meyerson, Q. Cronk, M. De Poorter, L.G. Eldrege, E. Green, M. Kairo, P. Latasi, R.N. Mack, J. Mauremootoo, D. O'Dowd, W. Orapa, S. Sastroutomo, A. Saunders, C. Shine, S. Thrainsson, and L. Vaiutu, 2007: Ecological and socioeconomic impacts of invasive alien species in island ecosystems. Environmental Conservation, 34(02), 98-111. Reenberg, A., T. Birch-Thomsen, O. Mertz, B. Fog, and S. Christiansen, 2008: Adaptation of human coping strategies in a small island society in the SW Pacific-50 Years of change in the coupled human-environment system on Bellona, Solomon Islands. Human Ecology, 36(6), 807-819. Restrepo, J.C., L. Otero, A.C. Casas, A. Henao, and J. Gutiérrez, 2012: Shoreline changes between 1954 and 2007 in the marine protected area of the Rosario Island Archipelago (Caribbean of Colombia). Ocean & Coastal Management, 69(0), 133-142. Richmond, N. and B.K. Sovacool, 2012: Bolstering resilience in the coconut kingdom: Improving adaptive capacity to climate change in Vanuatu. Energy Policy, 50, 843-848. Riedy, C. and I. McGregor, 2011: Climate Governance is Failing Us: We All Need to Respond. PORTAL Journal of Multidisciplinary International Studies, 8(3), 1449-2490. Rocha, S., I. Ineich, and J. Harris, 2009: Cryptic variation and recent bipolar range expansion within the Stumped- Toed Gecko Gehyra mutilata across Indian and Pacific Ocean islands. Contributions to Zoology, 78(1), 1-8. Rogers, T., K. Chmutina, and L.L. Moseley, 2012: The potential of PV installations in SIDS an example in the island of Barbados. Management of Environmental Quality: An International Journal, 23(3), 284-290. Romine, B.M. and C.H. Fletcher, 2013: A Summary of Historical Shoreline Changes on Beaches of Kauai, Oahu, and Maui, Hawaii. Journal of Coastal Research, 29(3), 605-614. Rongo, T. and R. van Woesik, 2011: Ciguatera poisoning in Rarotonga, southern Cook Islands. Harmful Algae, 10(4), 345-355. Rosenberg, E., A. Kushmaro, E. Kramarsky-Winter, E. Banin, and L. Yossi, 2009: The role of microorganisms in coral bleaching. The ISME Journal, 3(2), 139-146. Ross, M.S., J.J. O'Brien, R.G. Ford, K. Zhang, and A. Morkill, 2009: Disturbance and the rising tide: the challenge of biodiversity management on low-island ecosystems. Frontiers in Ecology and the Environment, 7(9), 471- 478. Rozell, D.J. and Wong T.F., 2010: Effects of climate change on groundwater resources at Shelter Island, New York State, USA. Hydrogeology Journal, 18(7), 1657-1665. Rudiak-Gould, P., 2012: Promiscuous corroboration and climate change translation: A case study from the Marshall Islands. Global Environmental Change, 22(1), 46-54. Russell, L., 2009: Poverty, Climate Change and Health in Pacific Island Countries. Menzies Centre for Health Policy, Sydney and Canberra, pp. 55. Santese, M., M.R. Perrone, A.S. Zakey, F. De Tomasi, and F. Giorgi, 2010: Modeling of Saharan dust outbreaks over the Mediterranean by RegCM3: Case studies. Atmospheric Chemistry and Physics Discussions, 10(1), 133- 156. Sax, D.F. and S.D. Gaines, 2008: Species invasions and extinction: the future of native biodiversity on islands. Proceedings of the National Academy of Sciences, 105, 11490-11497. Scheffers, A. and S. Scheffers, 2006: Documentation of the impact of Hurricane Ivan on the coastline of Bonaire (Netherlands Antilles). Journal of Coastal Research, 22(6), 1437-1450. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Schipper, L. and M. Pelling, 2006: Disaster risk, climate change and international development: scope for, and challenges to, integration. Disasters (Special Issue on Climate Change and Disasters), 30(1), 19-38. Schleupner, C., 2008: Evaluation of coastal squeeze and its consequences for the Caribbean island Martinique. Ocean & Coastal Management, 51(5), 383-390. Schofield, P.J., 2010: Update on geographic spread of invasive lionfishes (Pterois volitans [Linnaeus, 1758] and P. miles [Bennett, 1828]) in the Western North Atlantic Ocean, Caribbean Sea and Gulf of Mexico. Aquatic Invasions, 5(1), S117-S122. Schwarz, A., C. Béné, G. Bennett, D. Boso, Z. Hilly, C. Paul, R. Posala, S. Sibiti, and N. Andrew, 2011: Vulnerability and resilience of remote rural communities to shocks and global changes: Empirical analysis from Solomon Islands. Global Environmental Change, 21(3), 1128-1140. Scott, D., S. Gössling, and C.M. Hall, 2012a: International tourism and climate change. Wiley Interdisciplinary Reviews: Climate Change, 3(3), 213-232. Scott, D., S. Gössling, and C.M. Hall, 2012b: Climate Change and Tourism: Impacts, Adaptation and Mitigation. Routledge, London, pp. 437. Scott, D., M.C. Simpson, and R. Sim, 2012c: The vulnerability of Caribbean coastal tourism to scenarios of climate change related sea level rise. Journal of Sustainable Tourism, 20(6), 883-898. Sekercioglu, C.H., S.H. Schneider, J.P. Fay, and S.R. Loarie, 2008: Climate change, elevational range shifts, and bird extinctions. Conservation Biology, 22(1), 140-150. Senapathi, D., M.A.C. Nicoll, C. Teplitsky, C.G. Jones, and K. Norris, 2011: Climate change and the risks associated with delayed breeding in a tropical wild bird population. Proceedings of the Royal Society B: Biological Sciences, 278(1722), 3184-3190. Shen, S. and F. Gemenne, 2011: Contrasted Views on Environmental Change and Migration: the Case of Tuvaluan Migration to New Zealand. International Migration, 49(s1), e224-e242. Sinane, K., G. David, G. Pennober, and R. Troadec, 2010: Fragilisation et modification des formations littorales meubles sur l île d Anjouan (Comores): Quand l érosion d origine anthropique se conjugue au changement climatique. VertigO, La Revue Électronique En Sciences De l environnement, 10(3). Solomon, S.M. and D.L. Forbes, 1999: Coastal hazards and associated management issues on South Pacific Islands. Ocean & Coastal Management, 42(6), 523-554. Sovacool, B.K., 2012: Expert views of climate change adaptation in the Maldives. Climatic Change, 114(2), 295- 300. Spennemann, D.H.R., 1996: Nontraditional settlement patterns and typhoon hazard on contemporary Majuro Atoll, Republic of Marshall Islands. Environmental Management, 20(3), 337-348. Srinivasan, U.T., 2010: Economics of climate change: risk and responsibility by world region. Climate Policy, 10(3), 298-316. Steinberger, J.K., J. van Niel, and D. Bourg, 2009: Profiting from negawatts: Reducing absolute consumption and emissions through a performance-based energy economy. Energy Policy, 37(1), 361-370. Stephenson, T.S., A.A. Chen, and M.A. Taylor, 2008: Toward the development of prediction models for the primary Caribbean dry season. Theoretical and Applied Climatology, 92, 87-101. Storey, D. and S. Hunter, 2010: Kiribati: an environmental 'perfect storm'. Australian Geographer, 41(2), 167-181. Stuart, E.K., 2006: Energizing the island community: a review of policy standpoints for energy in small island states and territories. Sustainable Development, 14(2), 139-147. Swart, R. and F. Raes, 2007: Making integration of adaptation and mitigation work: mainstreaming into sustainable development policies. Climate Policy, 7(4), 288-303. Tabucanon, G.M. and B. Opeskin, 2011: The Resettlement of Nauruans in Australia: An early case of failed environmental migration. The Journal of Pacific History, 46(3), 337-356. Takasaki, Y., 2011: Do the commons help augment mutual insurance among the poor? World Development, 39(3), 429-438. Tanzil, J.T.I., B.E. Brown, and A.W. Tudhope, 2009: Decline in skeletal growth of the coral Porites lutea from the Andaman Sea, South Thailand between 1984 and 2005. Coral Reefs., 28(2), 519-528. Taylor, M.A., A. Centella, J. Charlery, I. Borrajero, A. Benzanilla, J. Campbell, R. Rivero, T.S. Stephenson, F. Whyte, and R. Watson, 2007: Glimpses of the Future: A briefing from the PRECIS Caribbean Climate Change Project. Caribbean Community Climate Change Centre, Belmopan, Belize, pp. 1-24. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Taylor, M.A., F.S. Whyte, T.S. Stephenson, and J.D. Campbell, 2013: Why dry? Investigating the future evolution of the Caribbean low level Jet to explain projected Caribbean drying International Journal of Climatology, 33(3), 784-792. Techera, E., 2008: Customary law and community based conservation of marine areas in Fiji. In: Beyond the Global Village. Environmental Challenges inspiring Global Citizenship. [Hillerbrand, R. and R. Karlsson(eds.)]. Interdisciplinary Press, Oxford, UK, pp. 107. Tegart, W.J.M., G.W. Sheldon, and D.C. Griffiths, 1990: Executive summary. In: Climate Change: The IPCC Impacts Assessment. Contribution of Working Group II to the First Assessment Report of the Intergovernmental Panel on Climate Change. [Tegart, W.J.M., G.W. Sheldon, and D.C. Griffiths(eds.)]. Australian Government Publishing Service, Canberra, Australia, pp. 1-5. Teles, F.R., 2011: Update on dengue in Africa. Dengue Bulletin, Dec; 35, 35-51. Terry, J.P. and A.C. Falkland, 2010: Responses of atoll freshwater lenses to storm-surge overwash in the Northern Cook Islands. Hydrogeology Journal, 18(3), 749-759. Terry, J.P., S. Garimella, and R.A. Kostaschuk, 2002: Rates of floodplain accretion in a tropical island river system impacted by cyclones and large floods. Geomorphology, 42(3 4), 171-182. Tester, P.A., R.L. Feldman, A.W. Nau, S.R. Kibler, and R. Wayne Litaker, 2010: Ciguatera fish poisoning and sea surface temperatures in the Caribbean Sea and the West Indies. Toxicon, 56(5), 698-710. Thiengo, S.C., A. Maldonado, E.M. Mota, E.J.L. Torres, R. Caldeira, O.S. Carvalho, A.P.M. Oliveira, R.O. Simoes, M.A. Fernandez, and R.M. Lanfredi, 2010: The giant African snail Achatina fulica as natural intermediate host of Angiostrongylus cantonensis in Pernambuco, northeast Brazil. Acta Tropica, 115(3), 194-199. Thomas, A. and R. Leichenko, 2011: Adaptation through insurance: lessons from the NFIP. International Journal of Climate Change Strategies and Management, 3(3), 250-263. Thomas-Hope, E. and A. Jardine-Comrie, 2007: Valuation of environmental resources for tourism in small island developing states: Implications for planning in Jamaica. International Development Planning Review, 29(1), 93- 112. Tkachenko, K.S., 2012: The northernmost coral frontier of the Maldives: The coral reefs of Ihavandippolu Atoll under long-term environmental change. Marine Environmental Research, 82(0), 40-48. Tompkins, E.L., M.C. Lemos, and E. Boyd, 2008: A less disastrous disaster: Managing response to climate-driven hazards in the Cayman Islands and NE Brazil. Global Environmental Change-Human and Policy Dimensions, 18(4), 736-745. Tompkins, E.L., L. Hurlston, and W. Poortinga, 2009: Foreignness as a constraint on learning: the impact of migrants on disaster resilience in small islands. Environmental Hazards, 8(4), 263-277. Tompkins, E.L., W.N. Adger, E. Boyd, S. Nicholson-Cole, K. Weatherhead, and N. Arnell, 2010: Observed adaptation to climate change: UK evidence of transition to a well-adapting society. Global Environmental Change, 20(4), 627-635. Tsyban, A., J.T. Everett, and J. Titus, 1990: World oceans and coastal zones. In: Climate Change: The IPCC Impacts Assessment. [McG. Tegart, W.J., G.W. Sheldon, and D.C. Griffiths(eds.)]. Australian Government Publishing Service, Canberra, pp. 6.1-6.28. Uno, I., K. Eguchi, K. Yumimoto, T. Takemura, A. Shimizu, M. Uematsu, Z. Liu, Z. Wang, Y. Hara, and N. Sugimoto, 2009: Asian dust transported one full cicuit around the globe. Nature Geoscience, 2, 557-560. Uyarra, M.C., I.M. Cote, J.A. Gill, R.R.T. Tinch, D. Viner, and A.R. Watkinson, 2005: Island-specific preferences of tourists for environmental features: implications of climate change for tourism-dependent states. Environmental Conservation, 32(1), 11-19. van Aalst, M.K., T. Cannon, and I. Burton, 2008: Community level adaptation to climate change: the potential role of participatory community risk assessment. Global Environmental Change, 18(1), 165-179. van Alphen, K., H.S. Kunz, and M.P. Hekkert, 2008: Policy measures to promote the widespread utilization of renewable energy technologies for electricity generation in the Maldives. Renewable and Sustainable Energy Reviews, 12(7), 1959-1973. van der Werf, G.R., D.C. Morton, R.S. DeFries, J.G. Olivier, P.S. Kasibhatla, R.B. Jackson, G.J. Collatz, and J.T. Randrsn, 2009: CO2 emissions from forest loss. Nature Geoscience, 2, 737-738. van Hooidonk, R., J. Maynard, and S. Planes, 2013: Temporary refugia for coral reefs in a warming world. Nature Climate Change, 3, 508-511. Van Kleef, E., H. Bambrick, and S. Hales, 2010: The geographic distribution of dengue fever and the potential influence of global climate change. TropIKA.Net, , 0-0. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 van Ledden, M., G. Vaughn, J. Lansen, F. Wiersma, and M. Amsterdam, 2009: Extreme wave event along the Guyana coastline in October 2005. Continental Shelf Research, 29(1), 352-361. Van Nostrand, J.M. and J.G. Nevius, 2011: Parametric Insurance: Using Objective Measures to Address the Impacts of Natural Disasters and Climate Change. Environmental Claims Journal, 23(3-4), 227-237. Vassie, J.M., P.L. Woodworth, and M.W. Holt, 2004: An example of North Atlantic deep-ocean swell impacting ascension and St. Helena Islands in the Central South Atlantic. Journal of Atmospheric and Oceanic Technology, 21(7), 1095-1103. Voccia, A., 2012: Climate change: what future for small, vulnerable states? International Journal of Sustainable Development & World Ecology, 19(2), 101-115. Walsh, K.J.E., K.L. McInnes, and J.L. McBride, 2012: Climate change impacts on tropical cyclones and extreme sea levels in the South Pacific A regional assessment. Global and Planetary Change, 80 81, 149-164. Warrick, O., 2009: Ethics and methods in research for community-based adaptation: Reflections from rural Vanuatu. Participatory Learning and Action, 60, 76-87. Waycott, M., L.J. McKenzie, J.E. Mellors, J.C. Ellison, M.T. Sheaves, C. Collier, A.M. Schwarz, A. Webb, J.E. Johnson, and C.E. Payri, 2011: Vulnerability of mangroves, seagrasses and intertidal flats in the tropical Pacific to climate change. In: Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change. [Bell, D., J.E. Johnson, and J.A. Hobday(eds.)]. Secretariat of the Pacific Community, Noumea, New Caledonia., pp. 297- 368. Waycott, M., C.M. Duarte, T.J.B. Carruthers, R.J. Orth, W.C. Dennison, S. Olyarnik, A. Calladine, J.W. Fourqurean, K.L. Heck, A.R. Hughes, G.A. Kendrick, W.J. Kenworthy, F.T. Short, and S.L. Williams, 2009: Accelerating loss of seagrasses across the globe threatens coastal ecosystems. Proceedings of the National Academy of Sciences, 106(30), 12377-12381. Webb, N., S. Coulson, I. Hodkinson, W. Block, J. BalE, and A. Strathdee, 1998: The effects of experimental temperature elevation on populations of cryptostigmatic mites in high Arctic soils Pedobiologia, 42(4), 298- 308. Webb, A.P., and P.S. Kench, 2010: The dynamic response of reef islands to sea-level rise: Evidence from multi- decadal analysis of island change in the central Pacific. Global and Planetary Change, 72(3), 234-246. Weil, E. and A. Cróquer, 2009: Spatial variability in distribution and prevalence of Caribbean scleractinian coral and octocoral diseases. I. Community-level analysis. Diseases of Aquatic Organisms, 83(3), 195-208. Weil, E. and C.S. Rogers, 2011: Coral reef diseases in the Atlantic-Caribbean. In: Coral Reefs: An Ecosystem in Transition. [Dubinsky, Z. and N. Stambler(eds.)]. Springer Netherlands, Dordrecht; Heidelberg; London; New York, pp. 465-491-491. Weir, T. and Z. Virani, 2011: Three linked risks for development in the Pacific Islands: Climate change, disasters and conflict. Climate and Development, 3(3), 193-208. Westphal, M., M. Browne, K. MacKinnon, and I. Noble, 2008: The link between international trade and the global distribution of invasive alien species. Biological Invasions, 10(4), 391-398. Wetzel, F.T., H. Beissmann, D.J. Penn, and W. Jetz, 2013: Vulnerability of terrestrial island vertebrates to projected sea level rise. Global Change Biology, 19(7), 2058-2070. Wheeler, D., 2011: Quantifying vulnerability to climate change: Implications for adaptation assistance. CGD Working Paper 240. Center for Global Development, Washington, DC, pp. 49. White, I., T. Falkland, P. Perez, A. Dray, T. Metutera, E. Metai, and M. Overmars, 2007: Challenges in freshwater management in low coral atolls. Journal of Cleaner Production, 15(16), 1522-1528. White, I. and T. Falkland, 2010: Management of freshwater lenses in small Pacific islands. Hydrogeology Journal, 18(1), 227-246. Whyte, F.S., M.A. Taylor, T.S. Stephenson, and J.D. Campbell, 2008: Features of the Caribbean low level jet. International Journal of Climatology, 28(1), 119-128. Williams, G.J., I.S. Knapp, J.E. Maragos, and S.K. Davy, 2010: Modelling patterns of coral bleaching at a remote Pacific atoll. Marine Pollution Bulletin, 60(9), 1467-1476. Willis, J.K. and J.A. Church, 2012: Regional sea-level projection. Science, 336(6081), 550-551. Winters, L.A. and P.M. Martins, 2004: When comparative advantage is not enough: business costs in small remote economies. World Trade Review, 3(3), 347-383. Woodroffe, C.D., 2008: Reef-island topography and the vulnerability of atolls to sea-level rise. Global and Planetary Change, 62(1-2), 77-96. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Yamamoto, L. and M. Esteban, 2010: Vanishing island states and sovereignty. Ocean & Coastal Management, 53(1), 1-9. Yamano, H., H. Kayanne, T. Yamaguchi, Y. Kuwahara, H. Yokoki, H. Shimazaki, and M. Chicamori, 2007: Atoll island vulnerability to flooding and inundation revealed by historical reconstruction: Fongafale Island, Funafuti Atoll, Tuvalu. Global and Planetary Change, 57(3-4), 407-416. Yates, M.L., G. Le Cozannet, M. Garcin, E. Salai, and P. Walker, 2013: Multidecadal Atoll Shoreline Change on Manihi and Manuae, French Polynesia. Journal of Coastal Research, 29(4), 870-882. Zahibo, N., E. Pelinovsky, T. Talipova, A. Rabinovich, A. Kurkin, and I. Nikolkina, 2007: Statistical analysis of cyclone hazard for Guadeloupe, Lesser Antilles. Atmospheric Research, 84(1), 13-29. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 29-1: Climate change projections for the medium (500-700 ppm CO2-e) RCP4.5 scenario for the main Small Islands regions. The table shows the 25th, 50th (median) and 75th percentiles for surface temperature and precipitation based on averages from 42 CMIP5 global models (adapted from WGI AR5 Table 14.1). Mean net regional sea level change is evaluated from 21 CMIP5 models and includes regional non-scenario components (adapted from WGI AR5 Figure 13-20). RCP4.5 Annual Projected Change for 2081-2100 compared to 1986-2005 Small Island Temperature (OC) Precipitation (%) Sea Level (m) Region 25% 50% 75% 25% 50% 75% Range Caribbean 1.2 1.4 1.9 -10 -5 -1 0.5 0.6 Mediterranean 2.0 2.3 2.7 -10 -6 -3 0.4 - 0.5 Northern 1.2 1.4 1.7 0 1 4 0.5 0.6 tropical Pacific Southern 1.1 1.2 1.5 0 2 4 0.5 0.6 tropical Pacific North 1.3 1.5 2.0 5 9 20 0.4 0.5 Indian Ocean West 1.2 1.4 1.8 0 2 5 0.5 0.6 Indian Ocean Table 29-2: Summary of projected percentage changes in tropical Pacific tuna catches by 2036 and 2100 relative to 1980-2000 and the estimated resulting percentage change to government revenue (after Bell et al., 2011). Year and SRES Scenario Tuna Fishery 2035 2100 B1/A2 B1 A2 Western fishery + 11% -0.2% -21% Skipjack tuna Eastern fishery +37% +43% +27% Western fishery -2% -12% -24% Bigeye tuna Eastern fishery +3% -4% -18% Total Skipjack tuna +19% +12% -7% Total Bigeye tuna +0.3% -9% -27% Total Federated States of 0.8 to 1.7% -0.9 to -1.9% Change to Micronesia Government Solomon Islands 0.01 to 0.16% -0.03 to 0.77% Revenue Kiribati +11 to 18.4% +7.2 to 12.0% (%) Tuvalu +3.7 to 9.2% +2.5 to 6.2% Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 29-3: Type of island in the Pacific region and implications for hydro-meteorological hazards (after Campbell, 2009). Island type and size Island elevation, slope, Implications for hazard rainfall Continental - Large - High elevations River flooding more likely to be a problem - High biodiversity - River flood plains than in other island types. In Papua New - Well developed soils - Orographic rainfall Guinea high elevations expose areas to frost (extreme during El Nino). Volcanic High Islands - Relatively small land area - Steep slopes Because of size few areas are not exposed to - Barrier reefs - Less well developed river tropical cyclones. Streams and rivers subject - Different stages of erosion systems to flash flooding. Barrier reefs may ameliorate - Orographic rainfall storm surge. Atolls - Very small land area - Very low elevations Exposed to storm surge, king tides and high - Small islets surround a lagoon - Convectional rainfall waves. Narrow resource base. Exposed to - Larger islets on windward side - No surface (fresh) water fresh water shortages and drought. Water - Shore platform on windward side - Ghyben-Herzberg problems may lead to health hazards. - No or minimal soil (freshwater) lens Raised Limestone Islands - Concave inner basin - Steep outer slopes Depending on height may be exposed to storm - Narrow coastal plains - Sharp karst topography surge. Exposed to fresh water shortages and - No or minimal soil - No surface water drought. Water problems may lead to health hazards. Table 29-4: Selected key risks and potential for adaptation for small islands from the present-day to the long-term. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 29-5: Top ten countries in the Asia-Pacific region based on absolute and relative physical exposure to storms and impact on GDP (between 1998 and 2009) (after ESCAP and UNISDR, 2010). Rank Absolute exposure Relative exposure (% of the Absolute GDP loss Loss (as a % of GDP) (millions affected) population affected) ($billions) 1 Japan (30.9) North Mariana Islands Japan (1,226.7) North Mariana Islands (58.2) (59.4) 2 Philippines (12.1) Niue (25.4) Rep. of Korea (35.6) Vanuatu (27.1) 3 China (11.1) Japan (24.2) China (28.5) Niue (24.9) 4 India (10.7) Philippines (23.6) Philippines (24.3) Fiji (24.1) 5 Bangladesh (7.5) Fiji (23.1) Hong Kong (13.3) Japan (23.9) 6 Rep. of Korea (2.4) Samoa (21.4) India (8.0) Philippines (23.9) 7 Myanmar (1.2) New Caledonia (20.7) Bangladesh (3.9) New Caledonia (22.4) 8 Viet Nam (0.8) Vanuatu (18.3) North Mariana Islands Samoa (19.2) (1.5) 9 Hong Kong (0.4) Tonga (18.1) Australia (0.8) Tonga (17.4) 10 Pakistan (0.3) Cook Islands (10.5) New Caledonia (0.7) Bangladesh (5.9) Note: Small islands are highlighted in bold Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 29-1: Representative tropical island typologies. From top-left: a young, active volcanic island (with altitudinal zones) and limited living perimeter reefs (purple zone at outer reef edge), through to an atoll (centre bottom) and raised limestone island (bottom right) dominated by ancient reef deposits (brown + white fleck). Atolls have limited, low-lying land areas but well developed reef/lagoon systems. Islands composed of continental rocks are not included in this figure, but see Table 29-3] Figure 29-2: A comparison of the degree of confidence in the detection of observed impacts of climate change on tropical small islands with the degree of confidence in attribution to climate change drivers at this time. For example, the blue symbol No. 2 (Coastal Systems), indicates there is very high confidence in both the detection of sea-level rise consistent with global means and its attribution to climate change drivers; whereas the red symbol No. 17 (Human Systems) indicates whilst detection of casualties and damage during extreme events is very high, there is presently low confidence in the attribution to climate change. It is important to note that low confidence in attribution frequently arises due to the limited research available on small island environments. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 29-3: Time series of RCP scenarios annual projected temperature and precipitation change relative to 1986- 2005 for six small islands regions (using regions defined in AR5 WG1, Annex 1: Atlas of Global and Regional Climate Projections).] Thin lines denote one ensemble member per model, thick lines the CMIP5 multi-model mean. On the right-hand side the 5th, 25th, 50th (median), 75th and 95th percentiles of the distribution of 20-yr mean changes are given for 2081 2100 in the four RCP scenarios. Note that the model ensemble averages in the figure are for grid points over wide areas and encompass many different climate change signals. To get projections for a specific location and time period use the maps in the Atlas or the online interactive version at but please note that in regions with small islands the models basically simulate the climate of the surrounding ocean and local conditions on land may differ. [Illustration to be redrawn to conform to IPCC publication specifications.] Figure 29-4: Tropical and extra-tropical cyclone impacts on the coasts of small islands. Four types of impacts are distinguished here, black arrows showing the connections between them, based on the existing literature. An example of the chain of impacts associated with two extra-tropical cyclones centred to the east of Japan is illustrated by the red arrows. Swell waves generated by these events in December 2008 reached islands in the southwest Pacific and caused extensive flooding (3) that impacted soil quality (8), freshwater resources (9), and damaged crops (10), buildings (15), and transport facilities (16) in the region (Example based on Hoeke et al., 2013). Examples of tropical cyclone impacts on small island coasts with reference 1. Society Islands, French Polynesia, February 2010 (Etienne, 2012); 2. Taveuni, Fiji, March 2010 (Etienne and Terry, 2012); 3. Cook Islands (de Scally, 2008); Society and Autral Islands, French Polynesia, February 2010 (Etienne, 2012); 4. Viti Levu, Fiji, March 1997 (Terry et al., 2002); 5. Society Islands, French Polynesia, February 2010 (Etienne, 2012); 6. Curacao, Bonaire, Netherlands Antilles, November 1999 (Scheffers and Scheffers, 2006); Hawaiian Islands (Fletcher et al., 2008); 7. Bay Islands, Honduras, October 1998 (Cahoon et al., 2003); 8. Marshall Islands, June 1905 (Spennemann, 1996); 9. Pukapuka atoll, Cook Islands, February 2005 (Terry and Falkland, 2010); 10. Vanuatu, February 2004 (Richmond and Sovacool, 2012); 11. 12. 13. Tuamotu Islands, French Polynesia, 1982-83 (Dupon, 1987); 14. Grenada, September 2004 (OECS, 2004); 15. Grenada, September 2004 (OECS, 2004); Tubuai, Austral Islands, French Polynesia, February 2010 (Etienne, 2012); 16. Vanuatu, February 2004 (Richmond and Sovacool, 2012); Guadeloupe Island, October 2008 (Dorville and Zahibo, 2010); 17. Bora Bora, Raiatea, Maupiti, Tahaa, Huahine, Society Islands, February 2010 (Etienne, 2012); 18. Vanuatu, February 2004 (Richmond and Sovacool, 2012); 19. Tuamotu, French Polynesia, 1982-83 (Dupon, 1987). Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 29 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Examples of extra-tropical cyclone impacts on small island coasts with reference 1. Maldives, April 1987 (Harangozo, 1992); 2. Maldives, January 1955 (Maniku, 1990); 3. Maldives, April 1987 (Harangozo, 1992); 9. Solomon Islands, December 2008 (Hoeke et al., 2013); 10. Chuck, Pohnpei, Kosrae, Federated States of Micronesia, December 2008 (Hoeke et al., 2013); 15. Majuro, Marshall Islands, November 1979 (Hoeke et al., 2013); 16. Coral Coast, Viti Levu, Fiji, May 2011 (Hoeke et al., 2013); 17. Majuro, Kwajalein, Arno, Marshall Islands, December 2008 (Hoeke et al., 2013); 18. Bismark Archipelago, Papua New Guinea, December 2008 (Hoeke et al., 2013). Figure 29-5. The impact of alternative climate change adaptation actions or policies. Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 30. The Ocean Coordinating Lead Authors Ove Hoegh-Guldberg (Australia), Rongshuo Cai (China) Lead Authors Peter G. Brewer (USA), Victoria J. Fabry (USA), Karim Hilmi (Morocco), Sukgeun Jung (Republic of Korea), Elvira Poloczanska (Australia), Svein Sundby (Norway) Contributing Authors Johann Bell (New Caledonia), Christopher J. Brown (Australia), Michael T. Burrows (UK), Long Cao (USA), Simon Donner (Canada), C. Mark Eakin (USA), Arne Eide (Norway), Benjamin Halpern (USA), Charles R. McClain (USA), Skip McKinnell (Canada), Mary O Connor (Canada), Camille Parmesan (USA), R. Ian Perry (Canada), Anthony J. Richardson (Australia), David Schoeman (Australia), Sergio Signorini (USA), William Skirving (USA / Australia), Dáithí Stone (Canada / South Africa / USA), William Sydeman (USA), Rui Zhang (China), Ruben van Hooidonk (USA) Review Editors Ly Omar (Senegal), Carol Turley (UK) Volunteer Chapter Scientists Jo Davy (New Zealand), Sarah Lee (USA) Contents Executive Summary 30.1. Introduction 30.1.1. Major Sub-Regions within the Ocean 30.1.2. Detection and Attribution of Climate Change and Ocean Acidification in Ocean Sub-Regions 30.2. Major Conclusions from Previous Assessments 30.3. Regional Changes and Projections of Future Ocean Conditions 30.3.1. Physical Changes 30.3.1.1. Heat Content and Temperature 30.3.1.2. Sea Level 30.3.1.3. Ocean Circulation, Surface Wind and Waves 30.3.1.4. Solar Insolation and Clouds 30.3.1.5. Storm Systems 30.3.1.6. Thermal Stratification 30.3.2. Chemical Changes 30.3.2.1. Surface Salinity 30.3.2.2. Ocean Acidification 30.3.2.3. Oxygen Concentration 30.4. Global Patterns in the Response of Marine Organisms to Climate Change and Ocean Acidification 30.5. Regional Impacts, Risks, and Vulnerabilities: Present and Future 30.5.1. High Latitude Spring Bloom Systems 30.5.1.1. Observed Changes and Potential Impacts 30.5.1.2. Key Risks and Vulnerabilities 30.5.2. Equatorial Upwelling Systems Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.5.2.1. Observed Changes and Potential Impacts 30.5.2.2. Key Risks and Vulnerabilities 30.5.3. Semi-Enclosed Seas 30.5.3.1. Observed Changes and Potential Impacts 30.5.3.2. Key Risks and Vulnerabilities 30.5.4. Coastal Boundary Systems 30.5.4.1. Observed Changes and Potential Impacts 30.5.4.2. Key Risks and Vulnerabilities 30.5.5. Eastern Boundary Upwelling Ecosystems 30.5.5.1. Observed Changes and Potential Impacts 30.5.5.2. Key Risks and Vulnerabilities 30.5.6. Sub-Tropical Gyres 30.5.6.1. Observed Changes and Potential Impacts 30.5.6.2. Key Risks and Vulnerabilities 30.5.7. Deep Sea (>1000 m) 30.5.7.1. Observed Changes and Potential Impacts 30.5.7.2. Key Risks and Vulnerabilities 30.5.8. Detection and Attribution of Climate Change Impacts with Confidence Levels 30.6. Sectorial Impacts, Adaptation, and Mitigation Responses 30.6.1. Natural Ecosystem Services 30.6.2. Economic Sectors 30.6.2.1. Fisheries and Aquaculture 30.6.2.2. Tourism 30.6.2.3. Shipping 30.6.2.4. Offshore Energy and Mineral Resource Extraction and Supply 30.6.3. Human Health 30.6.4. Ocean-based Mitigation 30.6.4.1. Deep Sea Carbon Sequestration 30.6.4.2. Offshore Renewable Energy 30.6.5. Maritime Security and Related Operations 30.7. Synthesis and Conclusions 30.7.1. Key Vulnerabilities and Risks 30.7.2. Global Frameworks for Decision Making 30.7.3. Emerging Issues, Data Gaps, and Research Needs 30.7.3.1. Changing Variability and Marine Impacts 30.7.3.2. Surface Wind, Storms, and Upwelling 30.7.3.3. Declining O2 Concentrations 30.7.3.4. Ocean Acidification 30.7.3.5. Net Primary Productivity 30.7.3.6. Movement of Marine Organisms and Ecosystems 30.7.3.7. Understanding Cumulative and Synergistic Impacts 30.7.3.8. Reorganization of Ecosystems and Food Webs 30.7.3.9. Socio-ecological Resilience References Frequently Asked Questions 30.1: Can we reverse the climate change impacts on the ocean? 30.2: How can we use non-climate factors to manage climate change impacts on the ocean? 30.3: Does slower warming mean less impact on plants and animals? 30.4: How will marine primary productivity change? 30.5: Will climate change cause dead zones in the ocean? Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Executive Summary The Ocean plays a central role in Earth s climate and has absorbed 93% of the extra energy from the enhanced greenhouse effect and approximately 30% of anthropogenic CO2 from the atmosphere. Regional responses are addressed here by dividing the Ocean into seven sub-regions: High Latitude Spring Bloom Systems (HLSBS), Eastern Boundary Upwelling Ecosystems (EBUE), Coastal Boundary Systems (CBS), Equatorial Upwelling Systems (EUS), Sub-Tropical Gyres (STG), Semi-Enclosed Seas (SES), and the Deep Sea (DS; >1000 m). The eighth region, Polar Seas, is dealt with by Chapter 28 [Figure 30-1; WGI 3.2.5, Box 3.1, 3.8]. Global average sea surface temperatures have increased since both the beginning of the 20th Century and the 1950s (virtually certain). The average sea surface temperature (SST) of the Indian, Atlantic and Pacific Oceans has increased by 0.65, 0.41 and 0.31°C respectively over the period 1950 2009 (very likely, p- value<0.05). Changes in the surface temperatures of the ocean basins are consistent with temperature trends simulated by ocean-atmosphere models with anthropogenic greenhouse gas forcing over the past century (high confidence). Sub-regions within the Ocean also show robust evidence of change, with the influence of long-term patterns of variability (e.g., Pacific Decadal Oscillation, PDO; Atlantic Multi-decadal Oscillation, AMO) contributing to variability at regional scales, and making changes due to climate change harder to distinguish and attribute [30.3.1, Figure 30-2e g, Table 30-1; WGI 2.4.2-3, 3.2 3.8, 10.3.4, 14]. Uptake of CO2 has decreased ocean pH (approximately 0.1 unit over 100 years), fundamentally changing ocean carbonate chemistry in all ocean sub-regions, particularly at high latitudes (high confidence). The current rate of ocean acidification is unprecedented within the last 65 Ma (high confidence) if not the last 300 Ma (medium confidence). Warming temperatures, declining pH and carbonate ion concentrations represent risks to the productivity of fisheries and aquaculture, and the security of regional livelihoods given the direct and indirect effects of these variables on physiological processes (e.g., skeleton formation, gas exchange, reproduction, growth, and neural function) and ecosystem processes (e.g., primary productivity, reef building, and erosion) (high confidence) [6.2, 6.3, 30.3.1, 30.3.2; 6.1.2; WGI 3.8.2, Box 3.2, 5.3.1]. Regional changes observed in winds, surface salinity, stratification, ocean currents, nutrient availability, and oxygen depth profile in many regions may be a result of anthropogenic greenhouse gas emissions (low to medium confidences). Marine organisms and ecosystems are likely to change in response to these regional changes, although evidence is limited and responses uncertain [30.3, 30.5; 6.2, 6.3; WGI 2.5, 2.7 3.3 3.8, 10.4.2, 10.4.4]. Most, if not all, of the Ocean will continue to warm and acidify, although the rates will vary regionally (high confidence). Differences between Representative Concentration Pathways (RCPs) are very likely to be minimal until 2040 (high confidence). Projected temperatures of the surface layers of the Ocean, however, diverge as the 21st Century unfolds and will be 1 3°C higher by 2100 under RCP8.5 than RCP2.6 across most ocean sub-regions. The projected changes in ocean temperature pose serious risks and vulnerabilities to ocean ecosystems and dependent human communities (robust evidence, high agreement, high confidence) [6.5, 30.3.1, 30.3.2, 30.7.1 Figure 30-2e g, Table 30-3]. Rapid changes in physical and chemical conditions within ocean sub-regions have already affected the distribution and abundance of marine organisms and ecosystems. Responses of species and ecosystems to climate change have been observed from every ocean sub-region (high confidence). Marine organisms are moving to higher latitudes consistent with warming trends (high confidence), with fish and zooplankton migrating at the fastest rates, particularly in HLSBS regions. Changes to sea temperature have also altered the phenology or timing of key life-history events such as plankton blooms, and migratory patterns and spawning in fish and invertebrates over recent decades (medium confidence). There is medium to high agreement that these changes pose significant uncertainties and risks to fisheries, aquaculture and other coastal activities. Ocean acidification maybe driving similar changes (low confidence), although there is limited evidence and low agreement at present. The associated risks will intensify as ocean warming and acidification continue [Box CC-MB, 30.4, 30.5, 6.3, 6.4]. Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Regional risks and vulnerabilities to ocean warming and acidification can be compounded by non-climate related stressors such as pollution, nutrient runoff from land, and over-exploitation of marine resources, as well as natural climate variability (high confidence). These influences confound the detection and attribution of the impacts of climate change and ocean acidification on ecosystems yet may also represent opportunities for reducing risks through management strategies aimed at reducing their influence, especially in CBS, SES, and HLSBS [30.1.2, 30.5, 5.3.4, 18.3.3 4]. Recent changes to wind and ocean mixing within the highly productive HLSBS, EBUE, and EUS are likely to influence energy transfer to higher trophic levels and microbial processes. There is, however, limited evidence and low agreement on the direction and magnitude of these changes and their relationship to ocean warming and acidification (low confidence). In cases where NPP increases or is not consumed (e.g., Benguela EBUE, low confidence), the increased transfer of organic carbon to deep regions can stimulate microbial respiration and reduce O2 levels (medium confidence). Oxygen concentrations are also declining in the tropical Pacific, Atlantic, and Indian Oceans (particularly EUS) due to reduced O2 solubility at higher temperatures, and changes in ocean ventilation and circulation [Box CC-PP, 30.3.1, 30.3.2, 30.5.5 6, 6.3.3; WGI 3.8.3]. Global warming will result in more frequent extreme events and greater associated risks to ocean ecosystems (high confidence). In some cases (e.g., mass coral bleaching and mortality), projected increases will eliminate ecosystems, and increase risks and vulnerabilities to coastal livelihoods and food security (e.g., CBS in SE Asia; SES, CBS, and STG in the Indo-Pacific) (medium to high confidence). Reducing stressors not related to climate change represents an opportunity to strengthen the ecological resilience within these regions, which may help them survive some projected changes in ocean temperature and chemistry [5.4, 30.6.1, 30.5.3 4, 30.5.6, Figure 30-4, Box CC-CR, IPCC 2012]. The highly productive HLSBS in the North-eastern Atlantic has changed in response to warming oceans (medium evidence, high agreement), with a range of consequences for fisheries. These ecosystems are responding to recent warming, with the greatest changes being observed since the late 1970s in the phenology, distribution and abundance of plankton assemblages, and the reorganization of fish assemblages (high confidence). There is medium confidence that these changes will have both positive and negative implications depending on the particular HLSBS fishery and the time frame [Box CC-MB, Box 6-1, 6.4.1.2, 30.5.1, 30.6.2.1]. EUS, which support highly productive fisheries off equatorial Africa and South America, have warmed over the past 60 years (Pacific EUS: 0.43oC, Atlantic EUS: 0.54oC; very likely, p-value<0.05). While warming is consistent with changes in upwelling intensity, there is low confidence in our understanding of how EUS will change, especially in how El Nino-Southern Oscillation (ENSO) and other patterns of variability will interact in a warmer world. The risk, however, of changes to upwelling increases with average global temperature, posing significant uncertainties for dependent ecosystems, communities, and fisheries [30.5.2; WGI 14.4,]. The surface waters of the SES show significant warming from 1982 and most CBS show significant warming since 1950. Warming of the Mediterranean has led to the recent spread of tropical species invading from the Atlantic and Indian Oceans. Projected warming increases the risk of greater thermal stratification in some regions, which can lead to reduced O2 ventilation and the formation of hypoxic zones, especially in the Baltic and Black Seas (medium confidence). In some CBS, such as the East China Sea and Gulf of Mexico, these changes are further influenced by the contribution of nutrients from coastal pollution contributing to the expansion of hypoxic (low O2) zones. These changes are likely to influence regional ecosystems as well as dependent industries such as fisheries and tourism, although there is low confidence in the understanding of potential changes and impacts [Table 30-1, 30.5.3, 5.3.4.3]. Coral reefs within CBS, SES, and STG are rapidly declining as result of local (i.e., coastal pollution, overexploitation), and climate change (high confidence). Elevated sea temperatures drive impacts such as mass coral bleaching and mortality (very high confidence), with an analysis of the CMIP5 ensemble projecting the loss of coral reefs from most sites globally by 2050 under mid to high rates of ocean warming (very likely) [Figure 30-10, 30.5.3 4, 30.5.6, Box CC-CR]. Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The productive EBUE and EUS involve upwelling waters that are naturally high in CO2 concentrations and low in pH, and hence are potentially vulnerable to ocean warming and acidification (medium confidence). There is limited evidence and low agreement, as to how upwelling systems are likely to change (low confidence). Declining O2 and shoaling of the aragonite saturation horizon through ocean acidification increases the risk of upwelling water being low in pH and O2 with impacts on coastal ecosystems and fisheries, as has been seen already (e.g., California Current EBUE). These risks and uncertainties are likely to involve significant challenges for fisheries and livelihoods along the west coasts of South America, Africa, and North America (low to medium confidence) [30.3.2.2, 30.5.2, 30.5.5, Box CC-UP, Box CC-PP]. Chlorophyll concentrations measured by satellites have decreased in the STG of the North Pacific, Indian and North Atlantic Oceans by 9%, 12% and 11%, respectively, over and above the inherent seasonal and interannual variability from 1998 2010 (high confidence; p-value<0.05). Significant warming over this period has resulted in increased water column stratification, reduced mixed layer depth and possibly declines in nutrient availability and ecosystem productivity (limited evidence, medium agreement). The short timeframe of these studies against well-established patterns of long-term variability lead to the conclusion that these changes are about as likely as not due to climate change [30.5.6, Table 30-1, Box CC-UP, 6.3.4]. The world s most abundant yet difficult to access habitat, the DS, is changing (limited evidence, medium agreement), with warming between 700 2000 m from 1957 2010 likely to involve a significant anthropogenic signal (medium confidence). Decreased primary productivity of surface waters (e.g., STG) is likely to reduce the availability of organic carbon to DS ecosystems. Understanding of the risks of climate change and ocean acidification to the DS is important given the size of the DS region but is limited (low confidence) [30.5.7, Figure 30-2; WGI 3.2.4, Figure 3.2, 3.9]. Changes to surface wind and waves, sea level, and storm intensity will increase the vulnerability of ocean- based industries such as shipping, energy and mineral extraction (medium confidence). Risks to equipment and people may be reduced through the design and use of ocean-based infrastructure, together with the evolution of policy (medium agreement). Risks and uncertainties will increase with further climate change. New opportunities as well as risks for shipping, energy and mineral extraction, and international issues over access and vulnerability, may accompany warming waters, particularly at high latitudes [30.3.1, 30.6.2, 28.2.2, 28.2.5, 28.3.4, 10.2.2, 10.4.4, IPCC 2012]. Changes to ocean temperature, chemistry and other factors are generating new challenges for fisheries, as well as benefits (high agreement). Climate change is a risk to the sustainability of capture fisheries and aquaculture development, adding to the threats of over-fishing and other non-climate stressors. In EUS and STG, shifts in the distribution and abundance of large pelagic fish stocks will have the potential to create winners and losers among island nations and economies. There has been a boost in fish stocks of high latitude fisheries in the HLSBS of the North Pacific and North Atlantic, partly as a result of 30 years of increase in temperature. This is very likely to continue, although some fish stocks will eventually decline. A number of practical adaptation options and supporting international policies can minimize the risks and maximize the opportunities [30.6, 30.7, 7.4.2, 29.4]. Adaptation strategies for ocean regions beyond coastal waters are generally poorly developed but will benefit from international legislation and expert networks, as well as marine spatial planning (high agreement). Fisheries and aquaculture industries with high-technology and/or large investments, as well as marine shipping and oil and gas industries, have high capacities for adaptation due to greater development of environmental monitoring, modeling and resource assessments. For smaller-scale fisheries and developing nations, building social resilience, alternative livelihoods, and occupational flexibility represent important strategies for reducing the vulnerability of ocean-dependent human communities. Building strategies that include climate forecasting and early-warning systems can reduce impacts of warming and ocean acidification in the short term. Overall, there is a strong need to develop ecosystem-based monitoring and adaptation strategies to mitigate rapidly growing risks and uncertainties to the coastal and oceanic industries, communities and nations (high agreement) [30.6, 7.3.2.4]. Significant opportunity exists within the Ocean and its sub-regions for reducing the CO2 flux to the atmosphere (limited evidence, medium agreement). Ecosystems such as mangroves, seagrass and salt marsh offer Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 important carbon storage and sequestration opportunities (e.g., Blue Carbon; limited evidence, medium agreement). Blue Carbon strategies can also be justified in terms of the ecosystem services provide by coastal vegetated habitats such as protection against coastal erosion and storm damage, and maintenance of habitats for fisheries species. Sequestration of anthropogenic CO2 into deep ocean areas still faces considerable hurdles with respect to the expense, legality and vulnerability of storage sites and infrastructure. There are also significant opportunities with the Ocean for the development of offshore renewable energy such as wind and tidal power [30.6.1, 30.6.4]. International frameworks for collaboration and decision-making are critically important for coordinating policy that will enable mitigation and adaptation by the Ocean sectors to global climate change (e.g., United Nations Convention on the Law of the Sea, UNCLOS). These international frameworks offer an opportunity to solve problems collectively, including improving fisheries management across national borders (e.g., reducing illegal, unreported and unregulated fishing, IUU), responding to extreme events, and strengthening international food security. Given the importance of the Ocean to all countries, there is a need for the international community to progress rapidly to a whole of ocean strategy for responding to the risks and challenges posed by anthropogenic ocean warming and acidification [30.7.2]. 30.1. Introduction The Ocean exerts a profound influence as part of the Earth, interacting with its atmosphere, cryosphere, land, and biosphere to produce planetary conditions. It also directly influences human welfare through the provision and transport of food and resources, as well as by providing cultural and economic benefits, and indirectly through the regulation of atmospheric gas content and the distribution of heat and water across the planet. Chapter 30 examines the extent to which regional changes to the Ocean can be accurately detected and attributed to anthropogenic climate change and ocean acidification, building on the conclusions of Chapter 6, which focuses on the marine physiological and ecological responses to climate change and ocean acidification. Detailed assessment of the role of recent physical and chemical changes within the Ocean to anthropogenic climate change is provided in WGI (particularly Chapters 2, 3, 13, and 14). In Chapter 30, impacts, risks, and vulnerabilities associated with climate change and ocean acidification are assessed for seven ocean sub-regions, and the expected consequences and adaptation options for key ocean-based sectors are discussed. Polar oceans (defined by the presence of sea ice in the north and by the Polar Front in the south) are considered in Chapter 28. While climate change affects coastal and low-lying sub-regions of multiple nations, detailed discussion of potential risks and consequences for these regions occurs in the relevant chapters of this report (e.g., WGII Chapters 5 and 29 as well as other regional sections). 30.1.1. Major Sub-Regions within the Ocean The Ocean represents a vast region that stretches from the high tide mark to the deepest oceanic trench (11,030 m) and occupies 71% of the earth's surface. The total volume of the Ocean is approximately 1.3 billion km3, with approximately 72% of this volume being below 1000 m (Deep Sea (DS), 30.5.7). There are considerable challenges in assessing the regional impacts of climate change on the Ocean. Devising an appropriate structure in order to explore the influence of climate change across the entire Ocean region and the broad diversity of life forms and habitats is challenging. [Longhurst, 1998] identified over 50 distinct ecological provinces in the Ocean, defined by physical characteristics and the structure and function of phytoplankton communities. Longhurst s scheme, however, yields far more sub-regions than could be sensibly discussed in the space allocated within AR5. Consequently, we have used comparable principles but have divided the non-polar ocean into seven larger sub-regions similar to Barber [1988]. We recognize that these sub-regions do not always match physical-chemical patterns or specific geographies, and that they interact strongly with terrestrial regions through weather systems and the exchange of materials. Different ocean sub-regions may also have substantially different primary productivities and fishery catch. Notably, over 80% of fishery catch is associated with three ocean sub-regions: Northern Hemisphere High Latitude Spring Bloom Systems (HLSBS), Coastal Boundary Systems (CBS), and Eastern Boundary Upwelling Ecosystems Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (EBUE; Table SM30-1, Figure 30-1). The DS (>1000m) is included as a separate category that overlaps with the six other ocean sub-regions dealt with in Chapter 30. [INSERT FIGURE 30-1 HERE Figure 30-1: (a) Separation of the world s non-polar oceans into seven major sub-regions (excluding the polar oceans, which are considered in Chapter 28). The chlorophyll-a signal measured by SeaWiFS and averaged over the period from Sep 4, 1997 to 30 Nov 2010 (NASA) is provides a proxy for differences in marine productivity (with the caveats provided in Box CC-PP). Ecosystem structure and functioning, as well as key oceanographic features provided the basis for separating the Ocean into the sub-regions shown. The map insert shows the distribution of Deep Sea (DS) habitat (>1000 m; Bathypelagic and Abyssopelagic habitats combined). Numbers refer to: 1 = High Latitude Spring Bloom Systems (HLSBS); 2 = Equatorial Upwelling Systems (EUS); 3 = Semi-Enclosed Seas (SES); 4 = Coastal Boundary Systems (CBS); 5 = Eastern Boundary Upwelling Ecosystems (EBUE); 6 = Sub- Tropical Gyres (STG); and 7 = DS (>1000 m). (b) Relationship between fish catch and area for each ocean sub- region is shown in (a). Red columns: average fish catch (as millions tons yr-1) for the period 1970 2006. Blue columns: area (millions km2). The four left-hand columns (sub-regions HLSBS-North, CBS, EBUE, and SES) cover 20 % of the world oceans area and deliver 80% of the world s fish catches. The values for the percent area of the Ocean, primary productivity, and fishery catch for the major sub-regions are listed in Table SM30-1.]. 30.1.2. Detection and Attribution of Climate Change and Ocean Acidification in Ocean Sub-Regions The central goal of Chapter 30 is to assess the recent literature on the Ocean as a region for changes that can be attributed to climate change and/or ocean acidification. Detailed assessments of recent physical and chemical changes in the Ocean are outlined in WGI Chapters 2, 3, 6, 10, 13, and 14 (AR5). The detection and attribution of climate change and ocean acidification on marine organisms and ecosystems is addressed in Chapter 6. Chapter 30 draws on these chapters to investigate regional changes in the physical, chemical, ecological, and socio-economic aspects of the Ocean and the extent to which they can be attributed to climate change and ocean acidification. Generally, successful attribution to climate change occurs when the full range of possible forcing factors is considered and those related to climate change are found to be the most probable explanation for the detected change in question [18.2.1.1]. Comparing detected changes with the expectations of well-established scientific evidence also plays a central role in the successful attribution of detected changes. We attempt to do this for seven sub-regions of the Ocean. There are a number of general limitations to the detection and attribution of impacts to climate change and ocean acidification that are discussed elsewhere [18.2.1] along with challenges [18.2.2]. Different approaches and best practice guidelines are discussed in WGI Chapters 10 and 18 as well as in several other places [Hegerl et al., 2007; Hegerl et al., 2010; Stott et al., 2010]. The fragmentary nature of ocean observing, structural uncertainty in model simulations, the influence of long-term variability, and confounding factors unrelated to climate change (e.g., pollution, introduced species, overexploitation of fisheries) represent major challenges [Halpern et al., 2008; Hoegh-Guldberg et al., 2011b; Parmesan et al., 2011]. Different factors may also interact synergistically or antagonistically with each other and climate change, further challenging the process of detection and attribution [Hegerl et al., 2007; Hegerl et al., 2010]. 30.2. Major Conclusions from Previous Assessments An integrated assessment of the impacts of climate change and ocean acidification on the Ocean as a region was not included in recent IPCC assessments, although a chapter devoted to the Ocean in the Second Assessment Report (SAR) did attempt to assess the impacts of projected regional and global climate changes on the oceans [Ittekkot et al., 1996]. The fact that assessments for ocean and coastal systems are spread throughout previous IPCC assessment reports reduces the opportunity for synthesizing the detection and attribution of climate change and ocean acidification across the physical, chemical, ecological, and socio-economic components of the Ocean and its sub- regions. The IPCC Fourth Assessment Report (AR4) concluded, however, that while terrestrial sub-regions are warming faster than the oceans, Observations since 1961 show that the average temperature of the global ocean has increased to depths of at least 3000 m and that the ocean has been taking up over 80% of the heat being added to the Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 climate system. AR4 also concluded that sea levels had risen due to the thermal expansion of the Ocean but recognized that our understanding of the dynamics of glaciers and ice sheets was too limited to assess their likelihood or provide a best estimate or an upper boundary for sea level rise (AR4, SPM). Changes to ocean temperature and density have been identified as having the potential to alter large-scale ocean circulation. AR4 concluded that with respect to the Meridional Overturning Circulation (MOC) it is very likely that up to the end of the 20th Century the MOC was changing significantly at interannual to decadal time scales (AR4, WGI Chapter 5, Box 5.1), despite limited evidence of a slowing MOC. According to AR4, sea-level rise over the last 100 150 years is probably contributing to coastal erosion in many places , including the east coast of the United States and the United Kingdom (AR4, WGII Chapter 1). The AR4 assessment was virtually certain that rising atmospheric CO2 had changed carbonate chemistry of the ocean (i.e., buffering capacity, carbonate and bicarbonate concentrations), and that a decrease in surface pH of 0.1 had occurred over the global ocean (calculated from the uptake of anthropogenic CO2 between 1750 and 1994 ([Sabine et al., 2004; Raven et al., 2005]; AR4, 5.4.2.3, WGI Table 7.3). Large-scale changes in ocean salinity were also observed from 1955 1998 and were characterized by a global freshening in sub-polar latitudes and salinification of shallower parts of the tropical and subtropical oceans . In this case, freshening was observed in the Pacific, with increased salinity being observed in the Atlantic and Indian Oceans (AR4, WGI 5.3.2 5.3.5). These changes in surface salinity were qualitatively consistent with expected changes to surface freshwater flux. Freshening of mid and high latitude waters together with increased salinity at low latitudes were seen as evidence of changes in precipitation and evaporation over the oceans . Substantial evidence indicated that changing ocean conditions have extensively influenced marine ecosystems (AR4, WGII Table 1.5). AR4 noted that there is an accumulating body of evidence to suggest that many marine ecosystems, including managed fisheries, are responding to changes in regional climate caused predominately by warming of air and sea surface temperatures (SST) and to a lesser extent by modification of precipitation regimes and wind patterns (AR4, WGII 1.3.4.2). Observed changes in marine ecosystems and managed fisheries reported within AR4 included: changes to plankton community structure and productivity, the phenology and biogeography of coastal species, intertidal communities on rocky shores and kelp forests, and the distribution of pathogens and invasive species. Changes were also observed in coral reefs (primarily increased mass coral bleaching and mortality), migratory patterns and trophic interactions of marine birds, reptiles, and mammals, as well as of a range of other marine organisms and ecosystems (AR4, WGII Table 1.5), although a separate exercise in detection and attribution of changes to climate change (as done for terrestrial studies) was not done as part of AR4. 30.3. Recent Changes and Projections of Future Ocean Conditions Evidence that increasing concentrations of atmospheric CO2 have resulted in the warming and acidification of the upper layers of the Ocean has strengthened since AR4. Understanding the full suite of physical and chemical changes to the Ocean is critical to the interpretation of the past and future responses of marine organisms and ecosystems, especially with respect to the implications for coastal and low-lying areas. 30.3.1. Physical Changes 30.3.1.1. Heat Content and Temperature The Ocean has absorbed 93% of the extra heat arising from the enhanced greenhouse effect (1971 2010), with most of the warming (64%) occurring in the upper (0 700 m) ocean (1971 2010; WGI Section 3.2.3, Figure 3.2, Box 3.1). It is virtually certain that global average sea surface temperatures (SST) have increased since the beginning of the 20th Century, with improvements and growth of data sets and archives, and the understanding of errors and biases since AR4 (WGI 2.4.2). It is virtually certain that the upper ocean (0 700m depth) has warmed from 1971 2010 (Figure 30-2a), while it is likely that the surface layers of the Ocean have warmed from the 1870s to 1971. Rates of increase in temperature are highest near the surface of the Ocean (>0.1°C decade-1 in the upper 75 m from 1971 to 2010) decreasing with depth (0.015°C decade-1 at 700 m; Figure 30-2b, c). It is very likely that the intensification of this warming near the surface has increased thermal stratification of the upper ocean by about 4% Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 between 0 200 m depth from 1971 2010 in all oceans north of 40°S. It is likely that the Ocean has warmed between 700 2000 m from 1957 2010, with the warming signal becoming less apparent or non-existent at deeper depths (WGI 3.2.1 3.2.3, Figures 3.1 3.2, Figure 3.9). These changes include a significant anthropogenic signal (virtually certain) [Gleckler et al., 2012; Pierce et al., 2012], with the surface waters of all three ocean basins warming at different rates, that all exceed those expected if there were no changes to greenhouse gas forcing over the past century (Figure 30-2e-g). In this respect, the observed record also falls within the range of historical model outputs that include observed increases in the concentration of greenhouse gases as opposed to models that do not (Figure 30-2 e-g). Data archives such as HadISST1.1 contain sea surface temperatures (SST) reconstructed from a range of sources, allowing an opportunity to explore mean monthly, gridded, global SST from 1870 to the present [Rayner et al., 2003]. We used the published HadISST1.1 data set (higher temporal and spatial resolution than HadSST3) to explore trends in historic SST within our sub-regions (Figure 30-1a; see definition of regions in Figure SM30-1 and Table SM30-2, column 1). The median SST for 1871 1995 from the Comprehensive Ocean-Atmosphere Data Set (COADS) were merged with data from the UK Met Office Marine Data Bank (MDB) to produce monthly globally- complete fields of SST on a 1° latitude-longitude SST grid from 1870 to date. The surface layers of the three ocean basins have warmed (p-value<0.05, very likely), with the Indian Ocean (0.11°C decade-1) warming faster than the Atlantic (0.07°C decade-1) and Pacific (0.05°C decade-1) Oceans (high confidence) (Table 30-1). This is consistent with the depth-averaged (0 700 m) temperature trend observed from 1971 2010 (Figure 30-2a). While some regions (e.g., North Pacific) did not show a clear warming trend, most regions showed either significant warming in the average temperature, or significant warming in either/or the warmest and coolest months of the year, over the period 1950 2009 (HadISST1.1 data, Table 30-1). Trends in SST show considerable sub-regional variability (Table 30-1, Figure 30-2a). Notably, the average temperature of most HLSBS did not increase significantly from 1950 2009 (except in the Indian Ocean; Table 30-1) yet the temperatures of the warmest month (North and South Atlantic, and South-eastern Pacific) and of the coolest month (North and South Atlantic, and South Pacific) showed significant upward trends over this period (p-value<0.05; Table 30-1). The two EUS warmed from 1950 2009 (Pacific EUS, 0.07°C decade-1 and Atlantic EUS, 0.09°C decade-1; Table 30- 1). The average monthly SST of the SES did not warm significantly, although the temperature of the coolest month increased significantly within the Baltic Sea (0.35°C decade-1 or 2.11°C from 1950 2009), as did the temperatures of the warmest months in the Black (0.14°C decade-1 or 0.83°C from 1950 2009), Mediterranean (0.11°C decade-1 or 0.66°C from 1950 2009) and Red (0.05°C decade-1 or 0.28°C from 1950 2009) Seas over the period 1950 2009 (very likely) (Table 30-1). Studies over shorter periods (e.g., 1982 2006, [Belkin, 2009]) report significant increases in average SST of the Baltic (1.35°C), Black (0.96°C), Red (0.74°C), and Mediterranean (0.71°C) Seas. Such studies are complicated by the influence of patterns of long-term variability and by the small size and land-locked nature of SES. Coastal Boundary Systems (except the Caribbean and Gulf of Mexico) all showed highly significant (p-value<0.05) warming (0.09°C 0.13°C.decade-1, Table 30-1). Among the EBUE, the Canary and Californian current regions exhibited a significant rate of change in the average SST (0.09°C decade-1 and 0.12°C decade-1, respectively; p-value<0.05), while the Benguela and Humboldt currents did not show significant temperature changes from 1950 2009 (p-value>0.05; Table 30-1). There was some variability between current systems in terms of the behavior of the coolest and warmest months. The temperature of the coolest month increased significantly from 1950 2009 in the case of the Benguela and California currents (0.06°C decade-1 and 0.12°C decade-1 respectively, p-value<0.05), while there was a significant increase in the temperature of the warmest month in the case of the Canary and Humboldt currents (0.11°C decade-1 and 0.10°C decade-1, respectively, Table 30-1). The average temperature of STG showed complex patterns with increasing temperatures (1950 2009) in the Indian, South Atlantic, and South Pacific Oceans (very likely) (0.11, 0.08, and 0.06°C decade-1, respectively; p-value<0.05), but not in the North Atlantic or North Pacific Ocean (p-value>0.05). These rates are half the value reported over shorter periods (e.g., 1998 2010, Table 1 in [Signorini and McClain, 2012] and based on NOAA_OI_SST_V2 data). Given the sensitivity of coral reefs to temperature ([Eakin et al., 2010; Strong et al., 2011; Lough, 2012]; Box CC-CR), trends in key coral reef regions were also examined using the World Resources Institute s Reefs at Risk Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 report (www.wri.org) to identify HadISST1.1 grid cells containing coral reefs (Figure 30-4b). Grouping the results into six major coral reef regions, we found that coral reef waters (with the notable exception of the Gulf of Mexico and Caribbean) have shown strong increases in average temperature (0.07 0.13°C decade-1) as well as the temperature of the coolest (0.07 0.14°C decade-1) and warmest months (very likely) (0.07 0.12°C decade-1; Table 30-1). These trends in temperature have resulted in an absolute increase in sea temperature of 0.44 0.79°C from 1950 2009. [INSERT FIGURE 30-2 HERE Figure 30-2: (a) Depth-averaged 0 700 m temperature trend for 1971 2010 (longitudinal versus latitude, colors and gray contours in °C per decade). (b) Zonally averaged temperature trends (latitude versus depth, colors and gray contours in °C per decade) for 1971 2010, with zonally averaged mean temperature over plotted (black contours in °C). (c) Globally-averaged temperature anomaly (Time versus depth, colors and grey contours in °C) relative to the 1971 2010 mean. (d) Globally-averaged temperature difference between the Ocean surface and 200 m depth (Black: annual values; red: five year running mean). Panels (a) (d) from WGI Figure 3.1. (e) (g) Observed and simulated variations in past and projected future annual average SST over three ocean basins (excluding regions within 300 km of the coast). The black line shows estimates from HadISST1.1 observational measurements. Shading denotes the 5 95 percentile range of climate model simulations driven with historical changes in anthropogenic and natural drivers (62 simulations), historical changes in natural drivers only (25), and the Representative Concentration Pathways: Dark Blue: RCP2.6; Light Blue: RCP4.5; Green: RCP6.0, and Red: RCP8.5). Data are anomalies from the 1986 2006 average of the HadISST1.1 data (for the HadISST1.1 time series) or of the corresponding historical all-forcing simulations. Further details are given in Box 21-2.] [INSERT FIGURE 30-3 HERE Figure 30-3: Velocity at which sea surface temperature (SST) isotherms shifted (km decade-1) over the period 1960 2009 calculated using HaDISST1.1, with arrows indicating the direction and magnitude of shifts. Velocity of climate change is obtained by dividing the temperature trend in °C decade-1 by the local spatial gradient °C km-1. The direction of movement of SST is denoted by the direction of the spatial gradient and the sign of the temperature trend: towards locally cooler areas with a local warming trend or towards locally warmer areas where temperatures are cooling. Adapted from [Burrows et al., 2011].] [INSERT TABLE 30-1 HERE Table 30-1: Regional changes in sea surface temperature (SST) over the period 1950 2009 using the Ocean regionalization specified in Figure 30-1a (for further detail of regions defined for analysis, see Figure SM30-1 and Table 30-2, column 1). A linear regression was fitted to the average of all 1×1 degree monthly SST data extracted from the HadISST1.1 data set [Rayner et al., 2003] for each sub-region over the period 1950 2009. All SST values less than -1.8oC, together with all SST pixels that were flagged as being sea ice, were reset to the freezing point of seawater (-1.8oC) to reflect the sea temperature under the ice. Separate analyses were also done to explore trends in the temperatures extracted from the coldest-ranked and the warmest-ranked month of each year (Table SM30-2). The table includes the slope of the regression (°C decade-1), the p-value for the slope being different from zero and the total change over 60 years (i.e., the slope of linear regression multiplied by 6 decades) for each category. The p- values that exceed 0.05 plus the associated slope and change values have a gray background, denoting the lower statistical confidence in the slope being different from zero (no slope). Note, changes with higher p-values may still describe informative trends although the level of confidence is lower that the slope is different from zero.] Given the essential role that temperature plays in the biology and ecology of marine organisms (Box CC-MB, 6.2, 6.3, [Pörtner, 2002; Poloczanska et al., 2013]), the speed of isotherm migration ultimately determines the speed at which populations must either move, adapt or acclimate to changing sea temperatures [Pörtner, 2002; Burrows et al., 2011; Hoegh-Guldberg, 2012]. Burrows et al. [2011] calculated the rate at which isotherms are migrating as the ratio of the rate of SST change (°C yr-1) to the spatial gradient of temperature (°C km-1) over the period 1960 2009 (Figure 30-3). While many of these temperature trajectories are towards the polar regions, some are not and are influenced by features such as coastlines. This analysis and others (e.g., North Atlantic, González-Taboada and Anadón [2012]) reveals that isotherms in the Ocean are moving at high velocities (up to 200 km decade-1), especially at low latitudes (high confidence) (Figure 30-3). Other sub-regions showed smaller velocities with contracting isotherms (cooling) in some areas (e.g., the Central and North Pacific, and Atlantic Oceans, Figure 30-3). There are Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 also changes in the timing of seasonal temperatures in both spring and fall/autumn [Burrows et al., 2011; Poloczanska et al., 2013] which, together with other variables (e.g., light, food availability, geography), are likely to affect biological processes such as the migration of species to higher latitudes, and the timing and synchrony of reproductive and other seasonal behaviors. Significant excursions of sea temperature above long-term summer temperature maxima (or below long-term temperature minima) significantly affect marine organisms and ecosystems [Hoegh-Guldberg, 1999; Bensoussan et al., 2010; Crisci et al., 2011; Harley, 2011]. Consequently, calculating heat stress as a function of exposure time and size of a particular temperature anomaly has proven useful in understanding recent changes to organisms and ecosystems (e.g., coral reefs and thermal anomalies, [Strong et al., 2011]). The total heat stress accumulated over the period 1981 2010 was calculated using the methodology of [Donner et al., 2007] and a reference climatology based on 1985 2000 in which the highest monthly SST was used to define the thermal threshold, above which accumulated thermal stress was calculated as exposure time multiplied by stress or Degree Heating Months (DHM) as the running total over four consecutive months. While most sub-regions of the Ocean experienced an accumulation of heat stress (relative to a climatology based on the period 1985 2000), equatorial and high latitude sub-regions in the Pacific and Atlantic Oceans have the greatest levels of the accumulated heat stress (Figure 30-4a). These are areas rich in thermally-sensitive coral reefs (Figure 30-4b, [Strong et al., 2011]). There was also a higher proportion of years that have had at least one stress event (DHM>1) in the last 30 years (1981 2010, Figure 30-4c) than in the preceding 30 years (1951 1980, Figure 30-4c, d. [INSERT FIGURE 30-4 HERE Figure 30-4: Recent changes in thermal stress calculated using HadISST1.1 data. A monthly climatology was created by averaging the HadISST monthly SST values over the period 1985 2000 to create twelve averages, one for each month of the year. The Maximum Monthly Mean (MMM) climatology was created by selecting the hottest month for each pixel. Anomalies were then created by subtracting this value from each SST value, but only allowing values to be recorded if they were greater than zero [Donner et al., 2007]. Two measures of the change in thermal stress were calculated as a result: (a) The total thermal stress for the period 1981 2010, calculated by summing all monthly thermal anomalies for each grid cell. (b) The location of coral reef grid cells used in Table 30-1 and for comparison to regional heat stress here. Each dot is positioned over a 1×1 degree grid cell within which lies at least one carbonate coral reef. The latitude and longitude of each reef is derived from data provided by the World Resources Institute s Reefs at Risk report (http://www.wri.org). The six regions are as follows: Red Western Pacific Ocean; Yellow Eastern Pacific Ocean; Dark Blue Caribbean & Gulf of Mexico; Green Western Indian Ocean; Pink Eastern Indian Ocean; and Light Blue Coral Triangle & SE Asia. (c) Proportion of years with thermal stress, which is defined as any year that has a thermal anomaly, for the periods 1951 1980 and (d) 1981 2010.] The three ocean basins will continue warming under moderate (RCP4.5) to high (RCP8.5) emission trajectories (high confidence) and will only stabilize over the second half of the century in the case of low range scenarios such as RCP2.6 (Figure 30-2 e-g; WGI, AI.4 AI.8). Projected changes were also examined for specific ocean sub-regions using ensemble averages from AOGCM simulations available in the CMIP5 archive (Table SM30-3) for the four scenarios of the future (Representative Concentration Pathways: RCP2.6, RCP4.5, RCP6.0 and RCP8.5; [van Vuuren et al., 2011]). Ensemble averages for each RCP are based on simulations from 10 16 individual models (Table SM30-3). The subset of CMIP5 models were chosen because each has historic runs enabling the derivation of the MMM climatology from 1985 2000, ensuring that all anomalies were comparable across time periods and across RCPs (Figure 30-10). Model hind-cast changes matched those observed for ocean sub-regions for the period 1980 2009 (HadISST1.1; Figure 30-2), with the model ensemble slightly overestimating the extent of change across the different ocean sub-regions (slope of observed/model = 0.81, r2 = 0.76, p-value<0.001). In this way, the absolute amount of change projected to occur in the ocean sub-regions was calculated for near-term (2010 2039) and long- term (2070 2099) periods (Table SM30-4). In the near-term, changes in the temperature projected for the surface layers of the Ocean are largely indistinguishable between the different RCP pathways due to the similarity in forcing until 2040. By the end of the century, however, SST across the ocean sub-regions were 1.8 3.3°C higher under RCP8.5 than those projected to occur under RCP2.6 (Table SM30-4; Figure 30-2 e g). The implications of these projected changes on the structure and function of oceanic systems are discussed below. Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.3.1.2. Sea Level The rate of sea level rise since the mid-19th Century has been larger than the mean rate during the previous two millennia (high confidence). Over the period 1901 2010, Global Mean Sea Level (GMSL) rose by 0.19 (0.17 0.21) m (WGI Figure SPM, 3.7, 5.6, 13.2). It is very likely that the mean rate of global averaged sea level rise was 1.7 [1.5 1.9] mm yr 1 between 1901 and 2010, 2.0 [1.7 2.3] mm yr 1 between 1971 and 2010 and 3.2 [2.8 3.6] mm yr 1 between 1993 and 2010 (WGI SPM, 3.7). These observations are consistent with thermal expansion of the Ocean due to warming plus the addition of water from loss of mass by melting glaciers and ice sheets. Current rates of sea level rise vary geographically, and can be higher or lower than the GMSL for several decades at time due to fluctuations in natural variability and ocean circulation (Figure 30-5). For example, rates of sea level rise are up to three times higher than the GMSL in the Western Pacific and South-east Asian region, and decreasing in many parts of the Eastern Pacific for the period 1993 2012 as measured by satellite altimetry (Figure 30-5; WGI 13.6.5). Sea level rise under increasing atmospheric greenhouse gas concentrations will continue for hundreds of years, with the extent and rate of the increase in GMSL being dependent on the emission scenario followed. Central to this analysis is the millennial-scale commitment to further sea level rise that is likely to arise from the loss of mass of the Greenland and Antarctic ice sheets (WGI 13.5.4, Figure 13.13). Sea level rise is very likely to increase during the 21st Century relative to the period 1971 2010 due to increased ocean warming and the continued contribution of water from loss of mass from glaciers and ice sheets. There is medium confidence that median sea level rise by 2081 2100 relative to 1986-2005 will be (5 95% range of process-based models): 0.44 m for RCP2.6, 0.53 m for RCP4.5, 0.55 m for RCP6.0, and 0.74 m for RCP8.5. Higher values of sea level rise are possible but are not backed by sufficient evidence to enable reliable estimates of the probability of specific outcomes. Many semi-empirical model projections of GMSL rise are higher than process-based model projections (up to about twice as large), but there is no consensus in the scientific community about their reliability and there is thus low confidence in their projections (WGI 13.5.2 3; Table 13.6, Figure 13.12). It is considered very likely that increases in sea level will result in greater levels of coastal flooding and more frequent extremes by 2050 (WGI 13.7.2; [IPCC, 2012]). It is about as likely as not that the frequency of the most intense storms will increase in some ocean basins, although there is medium agreement that the global frequency of tropical cyclones is likely to decrease or remain constant (WGI 14.6, 14.8). While understanding of associated risks is relatively undeveloped, coastal and low-lying areas, particularly in the southern Asia, Pacific Ocean and North Atlantic regions, face increased flood risk (5.3.3.2, 8.2.3.4, 9.3.4.4. Future impacts of sea level rise include increasing penetration of storm surges into coastal areas and changing patterns of shoreline erosion (5.3), as well as the inundation of coastal aquifers by saltwater (5.4.2.5, 29.3.2). Regionally, some natural ecosystems may reduce in extent (e.g., mangroves), although examples of habitat expansion have been reported [Brown et al., 2011]. Overall, changes to sea level are very likely to modify coastal ecosystems such as beaches, salt marshes, coral reefs and mangroves (5.4.2, Box CC-CR), especially where rates of sea level rise are highest (e.g., South-east Asia and the Western Pacific). [INSERT FIGURE 30-5 HERE Figure 30-5. Map of the rate of change in sea surface height (geocentric sea level) for the period 1993 2012 derived from satellite altimetry. Also shown are relative sea level changes (gray lines) from selected tide gauge stations for the period 1950 2012. For comparison, an estimate of global mean sea level change is shown (red lines) with each tide gauge time series. The relatively large short-term oscillations in local sea level (gray lines) are due to the natural climate variability and ocean circulation. For example, the large regular deviations at Pago Pago are associated with the El Nino-Southern Oscillation. Figure originally presented in WGI (FAQ 13.1, Figure 1). 30.3.1.3. Ocean Circulation, Surface Wind, and Waves Circulation of atmosphere and ocean (and their interactions) drives much of the chemical, physical, and biological characteristics of the Ocean, shaping phenomena such as ocean ventilation, coastal upwelling, primary production, Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and biogeochemical cycling. Critical factors for transporting nutrients from deep waters to the marine primary producers in the upper layers of the ocean include wind-driven mixing and upwelling. There has been a poleward movement of circulation features, including a widening of the tropical belt, contraction of the northern polar vortex, and a shift of storm tracks and jet streams to higher latitudes (medium confidence, WGI 2.7.5, 2.7.6, 2.7.8, Box 2.5). Long-term patterns of variability (years to decades) continue to prevent robust conclusions regarding long-term changes in atmospheric circulation and winds in many cases (WGI 2.7.5). There is high confidence, however, that the increase in northern mid-latitude westerly winds from the 1950s to the 1990s, and the weakening of the Pacific Walker circulation from the late 19th Century to the 1990s have been largely offset by recent changes (WGI 2.7.5, 2.7.8, Box 2.5). Wind stress has increased since the early 1980s over the Southern Ocean (medium confidence) (WGI 3.4.4), and tropical Pacific since 1990 (medium confidence), while zonal mean wind stress may have declined by 7% in the equatorial Pacific from 1862 1990 due to weakening of the tropical Walker circulation (medium confidence) (WGI 3.4.4; [Vecchi et al., 2006]). For example, it is very likely that the sub-tropical gyres of the major ocean basins have expanded and strengthened since 1993. However, the short-term nature of observing means that these changes are as likely as not to be due to decadal variability and/or due to longer term trends in wind forcing associated with climate change (WGI 3.6). Other evidence of changes in ocean circulation is limited to relatively short-term records that suffer from low temporal and spatial coverage. Therefore, there is very low confidence that multi-decadal trends in ocean circulation can be separated from decadal variability (WGI 3.6.6). There is no evidence of a long-term trend in large-scale currents such as the Atlantic Meridional Overturning Circulation (AMOC), Indonesian Throughflow (ITF), the Antarctic Circumpolar Current (ACC), or the transport of water between the Atlantic Ocean and Nordic Seas [WGI 3.6, Figures 3.10, 3.11]. Winds speed may have increased within the regions of EBUE (low confidence in attribution to climate; e.g. California Current, WGI 2.7.2). Changing wind regimes have the potential to influence mixed layer depth (MLD) and upwelling intensity in highly productive sub-regions of the world s oceans, although there is low agreement as to whether or not upwelling will intensify or not under rapid climate change ([Bakun, 1990; Bakun et al., 2010]; Box CC-UP). Surface waves are influenced by wind stress, although understanding trends remains a challenge due to limited data. There is medium confidence that Significant Wave Height (SWH) has increased since the mid-1950s over much of the North Atlantic north of 45°N, with typical winter season trends of up to 20 cm per decade (WGI 3.4.5). There is low confidence in the current understanding of how SWH will change over the coming decades and century for most of the Ocean. It remains an important knowledge gap (WGI 3.4). 30.3.1.4. Solar Insolation and Clouds Solar insolation plays a crucially important role in the biology of many marine organisms, not only as a source of energy for photosynthesis but also as a potential co-stressor in the photic zone (with temperature), as is seen during mass coral bleaching and mortality events (e.g., [Hoegh-Guldberg, 1999]). Global surface solar insolation (from the NCEP/NCAR Reanalysis Project, Kalnay et al. [1996]) decreased by 4.3 W m-2 decade-1 from the 1950s until 1991, after when it increased at 3.3 W m-2 decade-1 until 1999 [Ohmura, 2009; Wild, 2009], matching a broad suite of evidence from many land-based sites (WGI, 2.3.3). While there is consistency between independent data sets for particular regions, there is substantial ambiguity and therefore low confidence in observations of global-scale cloud variability and trends (WGI 2.5.7). There is also low confidence in projections of how cloudiness, solar insolation and precipitation will change as the planet warms due to the large interannual and decadal variability (ENSO, PDO), short observation time series and uneven spatial sampling, particularly in the early record (before 1950; WGI 2.5.8). 30.3.1.5. Storm Systems As agents of water column mixing, storms (from small atmospheric disturbances to intense tropical cyclones) can remix nutrients from deeper areas into the photic zone of the Ocean, stimulating productivity. Storms can also reduce local sea temperatures and associated stress by remixing heat into the deeper layers of the Ocean [Carrigan and Puotinen, 2011]. Large storms can destroy coastal infrastructure and coastal habitats such as coral reefs and mangrove forests, which can take decades to recover [Lotze et al., 2011; De ath et al., 2012]. While there is low Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 confidence for long-term trends in tropical cyclone activity globally (largely due to the lack of reliable long-term data sets), it is virtually certain that the frequency and intensity of the strongest tropical cyclones in the North Atlantic have increased since the 1970s (WGI 2.6.3). There is medium agreement that the frequency of the most intense cyclones in the Atlantic has increased since 1987 (WGI 2.6.3) and robust evidence of interdecadal changes in the storm track activity within the North Pacific and North Atlantic [Lee et al., 2012]. It is also very likely that there has been a decrease in the number of land-falling tropical cyclones along the East Australian coast since the 19th Century (WGI 2.6.3, [Callaghan and Power, 2011]). It is likely that these patterns are influenced by interannual variability such as ENSO, with land-falling tropical cyclones being twice as common in La Nina versus El Nino years (high confidence) [Callaghan and Power, 2011]. There has been an increase in the number of intense wintertime extra-tropical cyclone systems since the 1950s in the North Pacific. Similar trends have been reported for the Asian region, although analyzes are limited in terms of the spatial and temporal coverage of reliable records (WGI 2.6.4). There is low confidence, however, in large-scale trends in storminess or storminess proxies over the last century due to the lack of long-term data and inconsistencies between studies (WGI 2.6.4). 30.3.1.6. Thermal Stratification As heat has accumulated in the Ocean there has been a 4% increase in thermal stratification of the upper layers in most ocean regions (0 200 m, 40-year record) north of 40°S (WGI 3.2.2). Increasing thermal stratification has reduced ocean ventilation and the depth of mixing in many ocean sub-regions (medium confidence) WGI 3.8.3). This in turn reduces the availability of inorganic nutrients and consequently primary productivity (medium confidence) (6.3.4). In the STG, which dominate the three major ocean basins (30.5.6), satellite-derived estimates of surface chlorophyll and primary production decreased between 1999 and 2007 (Box CC-PP). In contrast, however, in situ observations at fixed stations in the North Pacific and North Atlantic Oceans (Hawaii Ocean Time-series or HOT, and Bermuda Atlantic Time-series Study, BATS), showed increases in nutrient and chlorophyll levels and primary production over the same period, suggesting that other processes (e.g., ENSO, PDO, NAO, winds, eddies, advection) can counteract broad-scale trends at local scales (Box CC-PP). The continued warming of the surface layers of the Ocean will very likely further enhance stratification and potentially limit the nutrient supply to the euphotic zone in some areas. The response of upwelling to global warming is likely to vary between regions and represents a complex interplay between local and global variables and processes (Box CC-UW). 30.3.2. Chemical Changes 30.3.2.1. Surface Salinity The global water cycle is dominated by evaporation and precipitation occurring over ocean regions, with surface ocean salinity varying with temperature, solar radiation, cloud cover, and ocean circulation [Deser et al., 2004]. Changes in salinity influence stratification of water masses and circulation. Ocean salinity varies regionally (Figure 30-6a) and is an outcome of the balance between evaporation and precipitation ([Durack and Wijffels, 2010]; WGI 3.3). Evaporation-dominated regions (Figure 30-6b) such as the STG, and Atlantic and Western Indian Oceans (WGI 3.3.3) have elevated salinity, while areas of high precipitation such as the North Pacific, North-eastern Indian Ocean, South-east Asia, and the eastern Pacific have relatively low salinities (WGI 3.3.3, Figure 30-6a). It is very likely that large-scale trends in salinity have also occurred in the Ocean interior, deriving from changes to salinity at the surface and subsequent subduction (WGI 3.3.2 3.3.4). Salinity trends are consistent with the amplification of the global hydrological cycle [Durack et al., 2012; Pierce et al., 2012], a consequence of a warmer atmosphere very likely producing the observed trend in greater precipitation, evaporation, atmospheric moisture (Figure 30-6b), and extreme events (WGI 2.6.2.1, 3.3.4; [IPCC, 2012]). Spatial patterns in salinity and evaporation-precipitation are similar, providing indirect evidence that these processes have been enhanced since the 1950s [WGI 3.3.2 3.3.4, Figures 3.4, 3.5 and 3.20d, FAQ 3.3]. These trends in salinity are very likely to have a discernible contribution from anthropogenic climate change (WGI 10.4.2). The combined changes in surface salinity and temperature are consistent with changes expected due to anthropogenic forcing of the climate system and are inconsistent with the effects of natural climate variability, either internal to the climate Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 system (e.g., ENSO, PDO; Figure 30-6c, d) or external to it (e.g., solar forcing or volcanic eruptions; [Pierce et al., 2012]). There is high confidence between climate models that the observed trends in ocean salinity will continue as average global temperature increases [Durack and Wijffels, 2010; Terray et al., 2012]. Ramifications of these changes are largely unknown but are of interest given the role of ocean salinity and temperature in fundamental processes such as the Atlantic Meridional Overturning Circulation (AMOC). [INSERT FIGURE 30-6 HERE Figure 30-6: (a) The 1955 2005 climatological-mean sea surface salinity [Antonov et al., 2010] color contoured at 0.5 PSS78 intervals (black lines). (b) Annual mean evaporation-precipitation averaged over the period 1950 2000 (NCEP) color contoured at 0.5 m yr 1 intervals (black lines). (c) The 58-year (2008 minus 1950) sea surface salinity change derived from the linear trend (PSS78), with seasonal and ENSO signals removed [Durack and Wijffels, 2010] color contoured at 0.116 PSS78 intervals (black lines). (d) The 30-year (2003 2007 average centered at 2005, minus the 1960 1989 average centered at 1975) sea surface salinity difference (PSS78) color contoured at 0.06 PSS78 intervals (black lines). Contour intervals in (c) and (d) are chosen so that the trends can be easily compared, given the different time intervals in the two analyzes. White areas in (c) and (d) are marginal seas where the calculations are not carried out. Regions where the change is not significant at the 99% confidence level are stippled in gray. Figure originally presented as WGI Figure 3.4 in WGI.] 30.3.2.2. Ocean Acidification The Ocean has absorbed approximately 30% of atmospheric CO2 from human activities resulting in decreased ocean pH and carbonate ion concentrations, and increased bicarbonate ion concentrations (Box CC-OA, WG1 Box 3.2; Figure SM30-2). The chemical response to increased CO2 dissolving into the Ocean from the atmosphere is known with very high confidence (WGI 6.4.4). Factors such as temperature, biological processes, and sea ice (WGI 6.4) play significant roles in determining the saturation state of seawater for polymorphs (i.e. different crystalline forms) of calcium carbonate. Consequently, pH and the solubility of aragonite and calcite are naturally lower at high latitudes and in upwelling areas (e.g., eastern Pacific upwelling, Californian Current ), where organisms and ecosystems may be relatively more exposed to ocean acidification as a result ([Feely et al., 2012; Gruber et al., 2012]; Figure 30-7a, b; Figure SM30-2). Aragonite and calcite concentrations vary with depth, with under-saturation occurring at deeper depths in the Atlantic (calcite: 3500 4500 m, aragonite: 400 3000 m) as opposed to the Pacific and Indian Oceans (calcite: 100 3000 m, aragonite: 100 1200 m; [Feely et al., 2004; Orr et al., 2005; Feely et al., 2009]; Figure 30-8). [INSERT FIGURE 30-7 HERE Figure 30-7: Projected ocean acidification from 11 CMIP5 Earth System models under RCP8.5 (other RCP scenarios have also been run with the CMIP5 models): (a) Time series of surface pH shown as the mean (solid line) and range of models (filled), given as area-weighted averages over the Arctic Ocean (green), the tropical oceans (red) and the Southern Ocean (blue). (b) Maps of the median model s change in surface pH from 1850 2100. Panel (a) also includes mean model results from RCP2.6 (dashed lines). Over most of the Ocean, gridded data products of carbonate system variables are used to correct each model for its present-day bias by subtracting the model-data difference at each grid cell following [Orr et al., 2005]. Where gridded data products are unavailable (Arctic Ocean, all marginal seas and the Ocean near Indonesia), the results are shown without bias correction. The bias correction reduces the range of model projections by up to a factor of 4, e.g., in panel (a) compare the large range of model projections for the Arctic (without bias correction) to the smaller range in the Southern Ocean (with bias correction). Figure originally presented in WGI Figure 6.28 in WGI.] [INSERT FIGURE 30-8 HERE Figure 30-8: Projected aragonite saturation state from 11 CMIP5 Earth System models under RCP8.5 scenario: (a) time series of surface carbonate ion concentration shown as the mean (solid line) and range of models (filled), given as area weighted averages over the Arctic Ocean (green), the tropical oceans (red), and the Southern Ocean (blue); maps of the median model's surface A in (b) 2010, (d) 2050, and (f) 2100; and zonal mean sections (latitude versus depth) of A in 2100 over (c) the Atlantic Ocean and (e) the Pacific Ocean, while the ASH (Aragonite Saturation Horizon) is shown for 2010 (dotted line) and 2100 (solid line). Panel (a) also includes mean model results from Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 RCP2.6 (dashed lines). As for Figure 30-7, gridded data products of carbonate system variables [Key et al., 2004] are used to correct each model for its present-day bias by subtracting the model-data difference at each grid cell following [Orr et al., 2005]. Where gridded data products are unavailable (Arctic Ocean, all marginal seas and the Ocean near Indonesia), results are shown without bias correction. Reprinted from Figure 6.29 in WGI.] Surface ocean pH has decreased by approximately 0.1 pH units since the beginning of the Industrial Revolution (high confidence) (Figure 30-7a; WGI 3.8.2, Box 3.2), with pH decreasing at the rate of -0.0013 and -0.0024 pH units yr-1 (WGI 3.8.2, Table 3.2). The presence of anthropogenic CO2 diminishes with depth. The saturation horizons of both polymorphs of calcium carbonate, however, are shoaling rapidly (1 2 m yr-1, and up to 5 m yr-1 in regions such as the California Current [Orr et al., 2005; Feely et al., 2012]. Further increases in atmospheric CO2 are virtually certain to further acidify the Ocean and change its carbonate chemistry (Figures S30.2, 30.7 and 30.8). Doubling atmospheric CO2 (~RCP4.5; [Rogelj et al., 2012]) will decrease ocean pH by another 0.1 units and decrease carbonate ion concentrations by approximately 100 umol kg-1 in tropical oceans (Figure 30-8a) from the present day average of 250 umol kg-1 (high confidence). Projected changes for the open Ocean by 2100 (Figures 30.7, 30.8) range from a pH change of -0.14 unit with RCP2.6 (421 ppm CO2, +1C, 22% reduction of carbonate ion concentration) to a pH change of -0.43 unit with RCP8.5 (936 ppm CO2, +3.7C, 56% reduction of carbonate ion concentration). The saturation horizons will also become significantly shallower in all oceans (with the aragonite saturation horizon between 0 and 1500 m in the Atlantic Ocean and 0 and 600 m (poles versus equator) in the Pacific Ocean ([Sabine et al., 2004; Orr et al., 2005]; WGI 6.4, Figure 6.28). Trends towards under-saturation of aragonite and calcite will also partly depend on ocean temperature, with surface polar waters expected to become seasonally under-saturated with respect to aragonite and calcite within a couple of decades (Figure 30-8c f, Box CC-OA[McNeil and Matear, 2008]). Overall, observations from a wide range of laboratory, mesocosm and field studies reveal that marine macro- organisms and ocean processes are sensitive to the levels of ocean acidification projected under elevated atmospheric CO2 (high confidence) (Box CC-OA, 6.3.2, [Munday et al., 2009; Kroeker et al., 2013]). Ecosystems that are characterized by high rates of calcium carbonate deposition (e.g., coral reefs, calcareous plankton communities) are sensitive to decreases in the saturation states of aragonite and calcite (high confidence). These changes are very likely to have broad consequences such as the loss of three-dimensional coral reef frameworks [Hoegh-Guldberg et al., 2007; Manzello et al., 2008; Fabricius et al., 2011; Andersson and Gledhill, 2013; Dove et al., 2013] and restructuring of food webs at relatively small (~50 ppm) additional increases in atmospheric CO2. Projected shoaling of the aragonite and calcite saturation horizons are likely to impact deep water (100 2000 m) communities of scleractinian corals and other benthic organisms as atmospheric CO2 increases ([Orr et al., 2005; Guinotte et al., 2006]; WGI 6.4), although studies from the Mediterranean and of seamounts off SW Australia report that some deep water corals may be less sensitive [Thresher et al., 2011; Maier et al., 2013]. Organisms are also sensitive to changes in pH with respect to physiological processes such as respiration and neural functions (6.3.2). Due to the relatively short history, yet growing effort, to understand the implications of rapid changes in pH and ocean carbonate chemistry, there are a growing number of organisms and processes reported to be sensitive. The impacts of ocean acidification on marine organisms and ecosystems continues to raise serious scientific concern, especially given that the current rate of ocean acidification (at least 10 100 faster than the recent glacial transitions [Caldeira and Wickett, 2003; Hoegh-Guldberg et al., 2007]) is unprecedented within the last 65 Ma (high confidence) [Ridgwell and Schmidt, 2010] and possibly 300 Ma of Earth history (medium confidence) ([Hönisch et al., 2012]; 6.1.2). 30.3.2.3. Oxygen Concentration Dissolved O2 is a major determinant of the distribution and abundance of marine organisms (6.3.3). Oxygen concentrations vary across ocean basins and are lower in the eastern Pacific and Atlantic basins, and northern Indian Ocean (Figure 30-9b, 6.1.1.3). In contrast, some of the highest concentrations of O2 are associated with cooler high latitude waters (Figure 30-9b). There is high agreement among analyzes providing medium confidence that O2 concentrations have decreased in the upper layers of the Ocean since the 1960s, particularly in the equatorial Pacific and Atlantic Oceans (WGI Figure 3.20, 3.8.3). A formal fingerprint analysis undertaken by Andrews et al. [2013] concluded that recent decreases in oceanic O2 are due to external influences (very likely). Conversely, O2 has Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 increased in the North and South Pacific, North Atlantic and Indian Oceans, consistent with greater mixing and ventilation due to strengthening wind systems (WGI 3.8.3). The reduction in O2 concentration in some areas of the Ocean is consistent with that expected from higher ocean temperatures and a reduction in mixing (increasing stratification) (WGI 3.8.3). Analysis of ocean O2 trends over time [Helm et al., 2011b] reveals that the decline in O2 solubility with increased temperature is responsible for no more than 15% of the observed change. The remaining 85%, consequently, is associated with increased deep-sea microbial respiration and reduced O2 supply due to increased ocean stratification (WGI Box 6.5 Figure 1). In coastal areas, eutrophication can lead to increased transport of organic carbon into adjacent ocean habitats where microbial metabolism is stimulated, resulting in a rapid drawdown of O2 [Weeks et al., 2002; Rabalais et al., 2009; Bakun et al., 2010]. The development of hypoxic conditions (generally defined as O2 concentrations below ~60 u mol kg-1) over recent decades has been documented across a wide array of ocean sub-regions including some SES (e.g., Black and Baltic Seas), the Arabian Sea, and the California, Humboldt, and Benguela Current systems, where eruptions of hypoxic, sulfide-laden water have also occurred in some cases [Weeks et al., 2002]. Localized, seasonal hypoxic dead zones have emerged in economically valuable coastal areas such as the Gulf of Mexico [Turner et al., 2008; Rabalais et al., 2010], the Baltic Sea [Conley et al., 2009] and the Black Sea [Kideys, 2002; Ukrainskii and Popov, 2009] in connection with nutrient fluxes from land. Over a vast region of the eastern Pacific stretching from southern Chile to the Aleutian Islands, the minimum O2 threshold (less than 2 mg l-1 or ~60 mmol-1) is found at 300 m depth and upwelling of increasingly hypoxic waters is well documented [Karstensen et al., 2008]. Hypoxic waters in the northern Arabian Sea and Bay of Bengal are located close to continental shelf areas. Long-term measurements reveal that O2 concentrations are declining in these waters, with medium evidence that economically significant mesopelagic fish populations are being threatened by a reduction in suitable habitat as respiratory stress increases [Koslow et al., 2011]. It should be noted that hypoxia profiles based on a critical threshold of 60 umol kg-1 can convey an overly simplistic message given that critical concentrations of O2 in this regard are very much species, size, temperature, and life history stage specific. This variability in sensitivity is, however, a critical determinant for any attempt to understand how ecosystems will respond to changing future O2 levels (6.3.3). There is high agreement among modeling studies that O2 concentrations will continue to decrease in most parts of the Ocean due to the effect of temperature on O2 solubility, microbial respiration rates, ocean ventilation, and ocean stratification (Figure 30-9c, d; WGI Table 6.14 [Andrews et al., 2013]), with implications for nutrient and carbon cycling, ocean productivity, marine habitats, and ecosystem structure (6.3.5). The outcomes of these global changes are very likely to be influenced by regional differences such as wind stress, coastal processes, and the supply of organic matter. [INSERT FIGURE 30-9 HERE Figure 30-9: (a) Simulated changes in dissolved O2 (mean and model range as shading) relative to 1990s for RCP2.6, RCP4.5, RCP6.0, and RCP8.5. (b) Multi-model mean dissolved O2 (mmol m 3) in the main thermocline (200 600 m depth average) for the 1990s, and changes in the 2090s relative to 1990s for RCP2.6 (c) and RCP8.5 (d). To indicate consistency in the sign of change, regions are stippled when at least 80% of models agree on the sign of the mean change. These diagnostics are detailed in [Cocco et al., 2013] in a previous model inter-comparison using the SRES- A2 scenario and have been applied to CMIP5 models here. Models used: CESM1-BGC, GFDL-ESM2G, GFDL- ESM2M, HadGEM2-ES, IPSL-CM5A-LR, IPSL-CM5A-MR, MPI-ESM-LR, MPI-ESM-MR, NorESM1. Figure originally presented in WGI Figure 6.30 in WGI.] 30.4. Global Patterns in the Response of Marine Organisms to Climate Change and Ocean Acidification Given the close relationship between organisms and ecosystems with the physical and chemical elements of the environment, changes are expected in the distribution and abundance of marine organisms in response to ocean warming and acidification (6.3, Box CC-MB, Box CC-OA). Our understanding of the relationship between ocean warming and acidification reveals that relatively small changes in temperature and other variables can result in often large biological responses that range from simple linear trends to more complex non-linear outcomes. There has been a rapid increase in studies that focus on the influence and consequences of climate change for marine ecosystems since AR4 ([Hoegh-Guldberg and Bruno, 2010; Poloczanska et al., 2013], representing an opportunity Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 to examine, and potentially attribute, detected changes within the Ocean to climate change. Evidence of global and regional responses of marine organisms to recent climate change have been shown through assessments of multiple studies focused on single-species, populations, and ecosystems [Tasker, 2008; Thackeray et al., 2010; Przeslawski et al., 2012; Poloczanska et al., 2013]. The most comprehensive assessment, in terms of geographic spread and number of observed responses, is that of Poloczanska et al. [2013]. This study reveals a coherent pattern in observed responses of ocean life to recent climate change across regions and taxonomic groups, with 81% of responses by organisms and ecosystems being consistent with expected changes to recent climate change (high confidence) (Box CC-MB). On average, spring events in the Ocean have advanced by 4.4 +/- 0.7 days decade-1 (mean +/- SE) and the leading edges of species distributions have extended (generally poleward) by 72.0 +/- 0.35 km decade-1. Values were calculated from data series ranging from the 1920s to 2010, although all series included data after 1990. The fastest range shifts generally occurred in regions of high thermal velocity (the speed and direction at which isotherms move [Burrows et al., 2011], 30.3.1.1). Subsequently, [Pinsky et al., 2013], using a database of 360 fish and invertebrate species and species groups from coastal waters around North America, showed differences in the speed and directions that species shift can be explained by differences in local climate velocities (Box CC-MB). 30.5. Regional Impacts, Risks, and Vulnerabilities: Present and Future This section explores the impacts, risks, and vulnerabilities of climate change for the seven sub-regions within the Ocean. There is considerable variability from region to region, especially in the extent and interaction of climate change and non-climate change stressors. While the latter may complicate attribution attempts in many sub-regions, interactions between the two groups of stressors may also represent opportunities to reduce the overall effects on marine organisms and processes by environmental changes being driven by climate change (including ocean acidification) [Crain et al., 2008; Griffith et al., 2012]. 30.5.1. High Latitude Spring Bloom Systems High Latitude Spring Bloom Systems (HLSBS) stretch from 35N to the edge of the winter sea ice (and from 35S to the polar front) and provide 36% of world s fish catch (Figure 30-1b). Although much of the North Pacific is iron limited [Martin and Fitzwater, 1988] and lacks a classical spring bloom [McAllister et al., 1960], strong seasonal variability of primary productivity is pronounced at all high latitudes because of seasonally varying photoperiod and water column stability [Racault et al., 2012]. Efficient transfer of marine primary and secondary production to higher trophic levels, including commercial fish species, is influenced by both the magnitude and the spatial and temporal synchrony between successive trophic production peaks [Hjort, 1914; Cushing, 1990; Beaugrand et al., 2003; Beaugrand and Reid, 2003]. 30.5.1.1. Observed Changes and Potential Impacts 30.5.1.1.1. North Atlantic The average temperature of the surface waters of the North Atlantic HLSBS has warmed by 0.07°C decade-1, resulting in an increase in sea temperature of 0.44°C between 1950 and 2009 (likely) (p-value = 0.15; Table 30-1). Over the same period, both winter and summer temperatures have increased significantly (0.05°C decade-1 and 0.12°C decade-1 respectively, p-value<0.05). Since the 1970s, the Atlantic Ocean has warmed more than any other ocean basin (0.3C decade-1; Figure 30-2a, WGI 3.2.2), with greatest warming rates over European continental shelf areas such as the southern North Sea, the Gulf Stream front, the sub-polar gyres and the Labrador Sea [MacKenzie and Schiedek, 2007b; a; Levitus et al., 2009; Lee et al., 2011; González-Taboada and Anadón, 2012]. Basin-wide warming in the North Atlantic since the mid-1990s has been driven by global warming and the current warm phase of the Atlantic Multidecadal Oscillation (AMO) ([Wang and Dong, 2010]; WGI 14.7.6 ). The North Atlantic is one of the most intensively fished ocean sub-regions. The major areas for harvesting marine living resources span the eastern North American, European and Icelandic shelves [Livingston and Tjelmeland, Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2000]. In addition, the deep regions of the Nordic Seas and the Irminger Sea contain large populations of pelagic fish such as herring, blue whiting and mackerel, and mesopelagic fish such as pearlsides and redfish. The region covers a wide latitudinal range from 35 80N and, hence, a large span in thermal habitats. This is reflected in the latitudinal gradient from subtropical/temperate species along the southern fringe to boreal/arctic species along the northern fringe. Climate change is virtually certain to drive major changes to the northern fringes of the Atlantic HLSBS by 2100. For the Barents Sea region, which borders the HLSBS and Arctic regions, modeling projections from 1995 2060 (SRES B2 scenario) gave an increase in phytoplankton production of 8%, an increase in Atlantic zooplankton production of 20%, and a decrease of Arctic zooplankton production of 50% [Ellingsen et al., 2008]. These changes result in a total increase in zooplankton production in the HLSBS section of the Barents Sea and a decrease in the Arctic section. Together with poleward shifts of fish species, a substantial increase in fish biomass and catch is also very likely at the northern fringes of the HLSBS [Cheung et al., 2011]. However, for some species like capelin, which feeds in summer at the ice edge and spawns in spring at the southern Atlantic Norwegian/Murman coast of the Barents Sea, the continuous temperature increase is very likely to cause discontinuous changes in conditions. The limited migration potential for this small pelagic fish is also likely to drive an eastwards shift in spawning areas to new spawning grounds along the Novaja Semlja coast[Huse and Ellingsen, 2008]. Observations of fish and other species moving to higher latitudes [Beare et al., 2005; Perry et al., 2005; Collie et al., 2008; Lucey and Nye, 2010] within the North Atlantic HLSBS are consistent with results of modeling exercises [Stenevik and Sundby, 2007; Cheung et al., 2011]. Examples from the Barents (28.2.2.1, Nordic, and North Seas (Box 6-1; 23.4.6) show how warming from the early 1980s influenced North Atlantic ecosystems, where substantial biological impacts such as large-scale modification of the phenology, abundance and distribution of plankton assemblages and reorganization of fish assemblages have been observed [Beaugrand et al., 2002; Edwards, 2004; Edwards and Richardson, 2004; Tasker, 2008; Nye et al., 2009; Head and Pepin, 2010; Simpson et al., 2011]. The ranges of some cold-water zooplankton assemblages in the North-east Atlantic have contracted towards the Arctic since 1958, and were replaced by warm-water zooplankton assemblages (specifically copepods) (high confidence), which moved up to 1000 km northward [Beaugrand et al., 2002; Beaugrand, 2009]. Although changes to surface circulation may have played a role [Reid et al., 2001], the primary driver of the shift was shown to be regional warming [Beaugrand et al., 2002; Beaugrand, 2004]. Reorganization of zooplankton communities and an observed decline in mean size has implications for energy transfer to higher trophic levels including commercial fish stocks ([Beaugrand et al., 2003; Kirby and Beaugrand, 2009; Lindley et al., 2010], 23.4.6). Warm-water species of fish have increased in abundance on both sides of the North Atlantic (medium confidence) [Beare et al., 2005; Collie et al., 2008; Genner et al., 2010; Hermant et al., 2010; Lucey and Nye, 2010; Simpson et al., 2011]. Diversity of zooplankton and fish has increased as more diverse warm-water assemblages extend northward in response to changing environmental conditions (high confidence) ([Kane, 2007; Hiddink and ter Hofstede, 2008; Beaugrand, 2009; Mountain and Kane, 2010; ter Hofstede et al., 2010], Box 6-1, 23.6.5). The past decade has been the warmest decade ever recorded in the Barents Sea, resulting in large populations of krill shrimp, and pelagic and demersal fish stocks linked to the Atlantic and boreal ecosystem of the Barents Sea (high confidence) ([Johannesen et al., 2012]; 28.2.2.1). Recruitment to boreal fish stocks such as cod, haddock, and herring has increased [Eriksen et al., 2012]. The relatively warm Atlantic waters have advanced northward and eastward [Arthun et al., 2012] and sea-ice has retreated along with the Arctic water masses. As a result, boreal euphausiids, which are mainly confined to Atlantic water, have increased in biomass and distribution [Dalpadado et al., 2012] enhancing growth of young cod Gadus morhua (boreal) as well as the more Arctic (arcto-boreal) capelin (Mallotus villosus). The abundance of amphipods of more Arctic origin has decreased, resulting in poorer feeding conditions for polar zooplankton predators such as polar cod (Boreogadus saida). Blue whiting (Micromesistius poutassou), which spawns west of the British Isles and feeds on zooplankton in the Norwegian Sea during the summer, extended their summer feeding distribution into the Barents Sea during the recent warm period. The Norwegian Sea is one of the two core regions for the herbivore copepod Calanus finmarchicus, an important prey species for pelagic fish and early life-stages of all fish around the rim of this high latitude sea including the North Sea and the Barents Sea [Sundby, 2000]. C. finmarchicus is the main food item for some of the world s largest fish stocks such as the Norwegian spring-spawning herring (Clupea harengus), blue whiting (M. poutassou), and Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 northeast Atlantic mackerel (Scomber scombrus). These stocks have increased considerably during the recent warming that started in the early 1980s [Huse et al., 2012]. The individual size of herring has also increased, enabling longer feeding migrations to utilize boreal zooplankton occurring closer to distant Arctic water masses. Mackerel (Scomber scombrus) has advanced northward and westward into Icelandic waters [Astthorsson et al., 2012] and was even observed in East Greenland water in summer 2013 [Nttestad et al., 2013]. Since 2004, the sum of spawning stock biomass of the three pelagic fish species (herring, blue whiting, and mackerel) leveled out at around 16 million tons. Observed changes in the phenology of plankton groups in the North Sea over the past 50 years are driven by climate forcing, in particular regional warming (high confidence) [Edwards and Richardson, 2004; Wiltshire and Manly, 2004; Wiltshire et al., 2008; Lindley et al., 2010; Lindley and Kirby, 2010; Schluter et al., 2010], although responses are species-specific with substantial variation within functional groups ([Edwards and Richardson, 2004]; Box 6-1). For example, the peak maximum abundance of the copepod C. finmarchicus advanced by 10 days from the 1960s to the 2000s, but its warm-water equivalent, C. helgolandicus, did not advance [Bonnet et al., 2005]. In the North Sea, bottom temperatures in winter have warmed by 1.6°C (1980 2004; [Dulvy et al., 2008]). The whole demersal fish community shifted deeper by 3.6 m decade-1 over the period 1980 2004, although mean latitude of the whole community did not show net displacement [Dulvy et al., 2008]. Within the community, cool-water specialists generally shifted northward while abundant warm-water species shifted southward, reflecting winter warming of the southern North Sea. The cold winter temperatures of the shallow regions of the southern North Sea have acted to exclude species with warm-water affinities. Trawl survey data from the rapidly-warming southern North Sea suggests waves of immigration by southern species such as red mullet (Mullus surmuletus), anchovy (Engraulis encrasicholus), and sardines (Sardina pilchardus), linked to increasing population sizes and warming temperatures [Beare et al., 2004; Beare et al., 2005]. In the North-east Atlantic, range expansions and contractions linked to changing climate have also been observed in benthic crustaceans, bivalves, gastropods, and polychaetes (medium confidence) [Mieszkowska et al., 2007; Beukema et al., 2009; Berke et al., 2010]. For example, the southern range limit of the common intertidal barnacle Semibalanus balanoides contracted northward along European coastlines at a rate of 15 50 km decade-1 since 1872, and its retreat is attributed to reproductive failure as winter temperatures warm [Southward et al., 2005; Wethey and Woodin, 2008]. Chthamalus montagui, its warm-water competitor, increased in abundance to occupy the niche vacated by S. balanoides (high confidence) [Southward et al., 1995; Poloczanska et al., 2008]. Many of the longest and most comprehensive time series used to investigate the ecological consequences of climate fluctuations and fishing, that span periods of cooling and warming over the past century, are from the North-east Atlantic [Toresen and Ostvedt, 2000; Southward et al., 2005; Sundby and Nakken, 2008; Edwards et al., 2010; Poloczanska et al., 2013]. Meta-analysis of 288 long-term datasets (spanning up to 90 years) of zooplankton, benthic invertebrates, fish, and seabirds from the OSPAR Commission Maritime Area in the North-east Atlantic showed widespread changes in distribution, abundance, and seasonality that were consistent (77%) with expectations from enhanced greenhouse warming [Tasker, 2008]. The study bought together evidence of changes in ocean climate and ecological responses across a range of species that encompassed both exploited and unexploited species from a variety of information types including peer-reviewed reports from International Council for the Exploration of the Sea (ICES) Working Groups. In particular, observations indicated poleward shifts in zooplankton communities, increasing abundance of fish species in the northern part of their ranges and decreases in southern parts, and the expansion of benthic species into more northerly or less coastal areas (high confidence). The major portion of the literature on the influence of climate change on the North Atlantic region covers time spans that are longer than for most other sub-regions of the Ocean. Even here, however, the bulk of the literature is limited to the last 30 50 years. The few publications covering the first half of the 20th Century represent an important longer-term perspective on the influence of climate change [Toresen and Ostvedt, 2000; Drinkwater, 2006; Sundby and Nakken, 2008; Banón, 2009; Astthorsson et al., 2012]. For example, distinct changes in fauna were associated with a pronounced warming period over 1920 1940 [Wood and Overland, 2010], when fish and other fauna shifted northward [Iversen, 1934; Southward et al., 2005; Drinkwater, 2006; Hátún et al., 2009]. The major lesson from these reports is that a rapid large-scale temperature increase occurred in the high latitude North Atlantic between the 1920s and 1940s, with basin-scale consequences for marine ecosystems that are comparable to warming and Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 observed impacts over the last 30 years. The former event was of great concern within the scientific community, particularly during the late 1940s and early 1950s [Iversen, 1934; Taning, 1949; Taning, 1953; Southward, 1980]. However, with the subsequent long-term cooling in the 1970s, discussion around climate responses was discontinued [Southward, 1980]. The centennial-long perspective indicates that multi-decadal variability has played a major role in changes observed over the past 30 years. The 150-yr instrumental record shows distinct warm phases of the Atlantic Multidecadal Oscillation (AMO) during approximately 1930 1965 and from 1995, and cool phases between approximately 1900 1930 and 1960 1995 (WGI 14.7.6). However, it is virtually certain that the enhanced warming in recent decades cannot be explained without external forcing (WGI 10.3.1.1.3). Understanding the changes in interdecadal variability over the next century is particularly important. The current warm phase of the AMO is likely to terminate in the next few decades, leading to a cooling influence in the North Atlantic and potentially offsetting some of the effects of global warming (WGI 14.7.6, 11.3.2.4.1). Over the transition period, the climate of the North Atlantic is likely to change more rapidly than during previous transitions since 1900. 30.5.1.1.2. North Pacific Sub-decadal variability in the North Pacific HLSBS is dominated by ENSO ([Trenberth, 1990]; WGI 14.4). Unlike the North Atlantic HLSBS, the North Pacific HLSBS does not show any significant trends in temperature over time, very likely as a consequence of climate variability influences on long-term warming patterns (1950 2009; Table 30- 1). Decadal and longer periods of variability in the North Pacific are reflected in the principal mode, the Pacific Decadal Oscillation (PDO; WGI 14.7.3), with periodicities in SST of both 15 25 y and 50 70 y [Minobe, 1997; Mantua and Hare, 2002]. Further modes of climate variability include the North Pacific Gyre Oscillation (NPGO; [Di Lorenzo et al., 2008; Chhak et al., 2009]. The PDO exhibits SST anomalies of one sign along the eastern boundary and the opposite sign in western and central Pacific. The PDO has been reported to have an anthropogenic component [Bonfils and Santer, 2011] but confidence in this is very low (limited evidence, low agreement) (WGI 10.3.3). The interplay of the phases of these modes of variability has strong influence on high latitude Pacific ecosystems (very high confidence). In the space of three years, the eastern North Pacific fluctuated from one of the warmest years in the past century (2005) to one of the coldest (2008) [McKinnell et al., 2010; McKinnell and Dagg, 2010]. This rapid change was accompanied by large changes in primary productivity, zooplankton communities, and fish and seabird populations [McKinnell et al., 2010; McKinnell and Dagg, 2010; Batten and Walne, 2011; Bi et al., 2011; Keister et al., 2011]. Climate transitions among phases of variability tend to be characterized by abrupt reorganization of the ecosystems as dynamic trophic relationships among species alter [Hunt et al., 2002; Peterson and Schwing, 2003; Litzow and Ciannelli, 2007; Litzow et al., 2008; Alheit, 2009]. Periods of broad-scale environmental change were observed across high latitude ecosystems in the North Pacific HLSBS (eastern Bering Sea and Gulf of Alaska) during 1976 78, 1987 89 and 1998 99. These periods were associated with regime shifts in foraging fish that occurred in 1979 82, 1988 92 and 1998 2001. The changes indicate how basin-scale variability such as the PDO can manifest across distinct ecosystems [Overland et al., 2008; Link et al., 2009a; Link et al., 2009b]. Phenological shifts observed in the zooplankton communities of the North Pacific were very likely in response to decadal climate variability, with distinct changes noted after the climate shifts of the 1970s and 1990s [Mackas et al., 1998; Peterson and Schwing, 2003; Chiba et al., 2006]. Modeling evidence suggests a weak shift in PDO towards more occurrences of the negative phase but credibility of projections remains uncertain (WGI 14.7.3). It is about as likely as not that the PDO will change its form or behavior in the future (WGI 14.7.3). The Kuroshio-Oyashio Extension (KOE) in the North-west Pacific displays pronounced decadal-scale variability [Yatsu et al., 2008; Sugisaki et al., 2010]. Warm periods in the mid-1970s and late-1980s were accompanied by dramatic changes in pelagic ecosystems and sardine and anchovy stocks [Chiba et al., 2008; Yatsu et al., 2008]. Observations and climate model simulations indicate that global warming is likely to further alter the dynamics of the Kuroshio Current and the KOE over the coming century [McPhaden and Zhang, 2002; Sakamoto et al., 2005; Wu et al., 2012; Zhang et al., 2013]. Alteration of the KOE will alter the timing, magnitude, and structure of spring blooms in the western Pacific and have implications for pelagic fish recruitment, production, and biogeochemical cycles [Ito et al., 2004; Hashioka et al., 2009; Yatsu et al., 2013]. Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Commercial catches of salmon species in the North Pacific HLSBS follow decadal fluctuations in climate [Hare and Mantua, 2000; Mantua and Hare, 2002]. Catches peaked in the warm periods of the 1930s 1940s and 1990s 2000s, with 2009 yielding the highest catch to date, and warming trends are about as likely as not to have contributed to recent peaks in some sub-regions [Morita et al., 2006; Irvine and Fukuwaka, 2011]. Poleward range shifts of some large pelagic fish in the western North Pacific, such as yellowtail Seriola quinqueradiata and Spanish mackerel Scomberomorus niphonius, were attributed, in part, to regional warming (high confidence) and these two species are projected to shift 39 71 km poleward from the 2000s to 2030s under SRES A1B [Tian et al., 2012; Jung et al., 2013]. Anticipating ecological responses to future anthropogenic climate change also requires evaluation of the role of changes to climate beyond warming per se. For example, declining sea level pressure (SLP) in the North Pacific is likely influenced by anthropogenic forcing [Gillett et al., 2003; Gillett and Stott, 2009] (WGI 10.3.3.4) and SLP in turn is related to atmospheric climate parameters (e.g., turbulent mixing via wind stress) that regulate commercially significant fish populations [Wilderbuer et al., 2002]. The northern fringe of the Bering Sea is among the most productive of marine sub-regions and includes the world s largest single-species fishery, walleye pollock Theragra chalcogramma [Hunt et al., 2010]. This region underwent major changes in recent decades as a result of climate variability, climate change, and fishing impacts ([Litzow et al., 2008; Mueter and Litzow, 2008; Jin et al., 2009; Hunt et al., 2010]; 28.2.2.1). Seasonal sea ice cover declined since the 1990s (to 2006), although there is no linear trend between 1953 and 2006, and the initiation of spring ice retreat over the south-eastern Bering Sea shelf became earlier [Wang et al., 2007a; Wang et al., 2007b]. Concurrent with the retreat of the cold pool , an area of reduced water temperature (<2°C) on the northern Bering Sea shelf that is formed as a consequence of sea ice and is maintained over summer [Hunt et al., 2010], bottom trawl surveys of fish and invertebrates show a significant community-wide northward distribution shift and a colonization of the former cold pool areas by sub-arctic fauna (high confidence) [Wang et al., 2006a; Mueter and Litzow, 2008]. Over a vast region of the eastern Pacific stretching from southern Chile to the Aleutian Islands, waters low in dissolved O2 (Oxygen Minimum Zone, OMZ) are found at 300 m depth [Karstensen et al., 2008]. Sporadic upwelling of these low-O2 waters along the continental shelf is well documented, where biological respiration can further reduce dissolved O2 levels and result in hypoxic or anoxic conditions that lead to mortality of coastal fishes and invertebrates [Grantham et al., 2004; Chan et al., 2008]. The magnitude and severity of seasonal hypoxic conditions in shallow-shelf waters of the eastern North Pacific HLSBS increased in recent decades [Bograd et al., 2008; Chan et al., 2008]. In addition, minimum pH values in the water column usually occur near the depths of the OMZ (WGI Box 3.2). A shoaling of the aragonite saturation horizon has likely resulted in low-aragonite conditions within the density layers being upwelled on the shelf of the west coast of the USA, increasing the risk of seasonally- upwelled water being relatively acidified [Feely et al., 2008] with observed impacts on Pacific oyster (Crassostrea gigas) hatcheries [Barton et al., 2012]. In the time period 1991 2006, reductions in pH in the North Pacific between 800 m and ~100 m were attributed in approximately equal measure to anthropogenic and natural variations ([Byrne et al., 2010]; WGI 3.8.2, Figure 3.19). 30.5.1.1.3. Southern Hemisphere The seasonal peaks in phytoplankton productivity in the southern hemisphere are much less pronounced and of smaller magnitude than those at northern hemisphere high latitudes [Yoder et al., 1993]. The southern hemisphere HLSBS is broadly bounded by the sub-tropical front (STF) and the sub-Antarctic front. Associated with the STF is intense biological activity of bloom-forming coccolithophores (phytoplankton) [Brown and Yoder, 1994]. The calcifying plankton assemblages play a key role in carbon cycles in the region and the transport of carbon to deep ocean sediments. The coccolithophore Emiliania huxleyi extended its range south of 60° in the South-west Pacific (141 145°E) over the two decades since 1983 [Cubillos et al., 2007]. Although the drivers for this range extension are not clear, it was proposed that the extension is facilitated by surface warming or changes in the abundance of grazing zooplankton. Large regions of the sub-Antarctic and Arctic surface waters are likely to become undersaturated with respect to aragonite during winter by 2030, which will impact calcifying plankton and Southern Ocean ecosystems ([McNeil and Matear, 2008; Bednar¹ek et al., 2012]; 28.2.2.2). Shell weights of the modern foraminifer Globigerina bulloides Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 in the sediments of the sub-Antarctic region of the HLSBS south of Australia were observed to be 30 35% lower than those from sediment cores representing pre-industrial periods, consistent with a recent decline in pH [Moy et al., 2009]. Examination of the pteropod Limacina helicina antarctica captured from polar waters further south, show severe levels of shell dissolution consistent with the shoaling of the aragonite saturation horizon and indicate that the impact of ocean acidification is already occurring [Bednar¹ek et al., 2012]. While the South Pacific HLSBS has not shown warming overall, both the warmest and coolest months show a slight, but significant, increase over time (both 0.05°C decade-1 from 1950 2009, p-value < 0.05, Table 30-1), although some areas within this sub-region have warmed. For example, the western Tasman Sea has shown enhanced warming since 1900 as compared to average global trends (high confidence). This has been driven by changes in large-scale wind-forcing leading to a southward expansion of the South Pacific STG and intensification of the southward-flowing East Australian Current (EAC; [Cai, 2006; Hill et al., 2008; Wu et al., 2012]; WG1 3.6.2). Model simulations suggest both stratospheric ozone depletion and greenhouse forcing contribute to the observed trend in wind stress [Cai and Cowan, 2007]. Coinciding with this warming and intensified EAC is the observation that a number of benthic invertebrates, fish, and zooplankton are now found further south than they were in the mid- 20th Century [Ling, 2008; Pitt et al., 2010; Last et al., 2011]. Warming facilitated the establishment of the grazing urchin Centrostephanus rodgersii in eastern Tasmania during the late 1970s (high confidence), which has resulted in deleterious effects on macroalgal beds [Ling, 2008; Ling et al., 2008; Ling et al., 2009; Banks et al., 2010]. 30.5.1.2. Key Risks and Vulnerabilities Projected changes to the temperature of surface waters match those of the past 50 years, with average sea temperatures in the HLSBS regions projected to increase by 0.35 1.17°C in the near-term (2010 2039) and by 1.70 4.84°C over the long-term (2010 2099) under the Business as usual (BAU) RCP8.5 scenario (Table SM30- 4). Under the lower-case scenario considered here (RCP2.6), projected rates of regional warming are much lower (0.12-0.79°C) in the near-term, with slight cooling for some regions in the long-term (-0.16 1.46°C). Risks to HLSBS from warming of surface waters include changes to primary production and carbon cycling, and the reorganization of ecosystems in response to warmer and more acidified oceans. Both primary production and the timing of the spring bloom in HLSBS are very sensitive to environmental change. Latitudinal shifts in the distribution of phyto- and zooplankton communities will alter seasonality, community composition, and bloom dynamics [Beaugrand, 2009; Ito et al., 2010; Shoji et al., 2011]. Alteration of the structure and composition of plankton communities can propagate through high latitude food webs due to tight trophic linkages [Edwards and Richardson, 2004; Beaugrand et al., 2010; Beaugrand and Kirby, 2010]. Mechanisms are complex, and tend to be non-linear, with impacts on ecosystems, fisheries, and biogeochemical cycles being hard to project with any certainty (Box CC-PP). A reorganization of commercial fish stocks, with attendant social and economic disruption, is a key risk of ongoing climate change in HLSBS sub-regions. AR4 reported that the productivity of some marine fisheries is likely to increase in the North Atlantic (AR4 WGII 10.4.1, 12.4.7). A large number of publications since then has substantially extended such documentation and begun to elucidate the nuances in how marine ecosystems and organisms respond [Sumaila et al., 2011]. An additional risk exists for sub-polar areas from the loss of seasonal sea ice. Decreases in seasonal sea ice are very likely to lead to increases in the length of the growth season and the intensity of the light available to fuel phytoplankton growth and, hence, enhance primary production and attending modifications of ecosystem structure [Arrigo et al., 2008]. In the long-term, however, primary production may decrease due to the reduced supply of nutrients to the surface layers (Box CC-PP). The decline in Arctic sea ice will open ecological dispersal pathways, as well as new shipping routes (30.6.2.3), between the North Atlantic and the North Pacific; large numbers of the Pacific diatom Neodenticula seminae were found in the North Atlantic in 1999 [Reid et al., 2007]. HLSBS are also vulnerable to rapid changes in the carbonate chemistry of ocean waters. Ocean acidification will produce additional and large-scale challenges. There is medium agreement that calcifying organisms in these regions will be negatively affected by ocean acidification with substantial impacts on higher trophic levels, although there is limited evidence at this point. Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.5.2. Equatorial Upwelling Systems The largest upwelling systems are found in the equatorial regions of the eastern Pacific and Atlantic Oceans (Figure 30-1a). Equatorial Upwelling Systems (EUS) produce highly productive cold tongues that stretch westward across equatorial areas, which is different to other upwelling systems (e.g., EBUE; 30.5.5). The associated upwelling is a consequence of the Earth s rotation and easterly (westward) winds and currents, which drive water northwards and southwards at the northern and southern edges of these sub-regions. As result, cold, nutrient-rich, and high CO2/low pH waters are transported from the deeper layers of the Ocean to the surface, driving high levels of primary productivity that support 4.7% of total global fisheries productivity (Table SM30-1, Figure 30-1b). Interannual modes of variability (e.g., ENSO; WGI 14.4) dominate EUS, particularly in the Pacific [Barber et al., 1994; McCarthy et al., 1996; Signorini et al., 1999; Le Borgne et al., 2002; Christian and Murtugudde, 2003; Mestas- Nunez and Miller, 2006; Pennington et al., 2006; Wang et al., 2006b]. Upwelling of the Pacific EUS declines during El Nino events, when the trade winds weaken, or even reverse, and is strengthened during La Nina events. ENSO periodicity controls primary productivity and consequently has a strong influence over associated fisheries production [Mestas-Nunez and Miller, 2006]. The Intertropical Convergence Zone (ITCZ; WGI 14.3.1.1), an important determinant of regional ocean temperature, is located at the edges of the Indian and Pacific equatorial upwelling zone and influences a range of variables including productivity, fisheries, and precipitation. The EUS are also affected by interdecadal variability (e.g., Interdecadal Pacific Oscillation (IPO); [Power et al., 1999]; WGI 11.2.2, 14.3). 30.5.2.1. Observed Changes and Potential Impacts The average sea temperature associated with the EUS has increased significantly (p-value<0.05) by 0.43°C and 0.54°C from 1950 2009 in the Pacific and Atlantic EUS, respectively [Table 30-1). In the Pacific, regional variability in SST trends is driven by the temporal patterns in El Nino-Southern Oscillation and the more frequent El Nino Modoki or Central Pacific El Nino events in recent decades (high confidence) ([Ashok et al., 2007; Yu and Kao, 2007; Lee and McPhaden, 2010]; WGI 14.2.4.4). The faster warming of the Atlantic EUS is likely to be associated with a weakening of upwelling [Tokinaga and Xie, 2011]. Sea level rise in the eastern equatorial Pacific has been decreasing by up to 10 mm yr-1 since 1993 ([Church et al., 2006]; Figure 30-5). Coral reefs in the EUS of the eastern Pacific (e.g., Galápagos and Cocos islands) have relatively low species diversity and poorly developed carbonate reef frameworks, due to the low pH and aragonite saturation of upwelling waters (high confidence) [Glynn, 2001; Manzello et al., 2008; Manzello, 2010]. Prolonged periods of elevated temperature associated with El Nino have negatively affected corals, kelps and associated organisms, and induced several possible local extinctions (high confidence) [Glynn, 2011]. Since 1985, coral reefs from west of South America to the Gilbert Islands of Kiribati have experienced the highest levels of thermal stress relative to other areas [Donner et al., 2010]. In 1982/1983, mass coral bleaching and mortality affected most of the reef systems within the eastern equatorial Pacific [Glynn, 1984; Baker et al., 2008]. Subsequent canonical El Nino and Central Pacific El Nino events in 1997/8, 2002/3, 2004/5, and 2009/10 (WGI 14.4.2, Figure 14.13) triggered mass coral bleaching by adding to the background increases in sea temperatures (high confidence) [Donner et al., 2010; Obura and Mangubhai, 2011; Vargas-Ángel et al., 2011]. In some locations, impacts of El Nino have also interacted with other anthropogenic changes, such as those arising from changes to fishing pressure [Edgar et al., 2010], further complicating the attribution of recent ecological changes to climate change. 30.5.2.2. Key Risks and Vulnerabilities Climate models indicate that ENSO is virtually certain to continue to be a major driver of oceanic variability over the coming century, although not all models can accurately replicate its behavior (WGI 9.5.3). Superposition of a warming ocean on future ENSO activity (possibly modified in frequency and intensity) is likely to result in oceanic conditions that are different from those experienced during past El Nino and La Nina events [Power and Smith, 2007]. Temperatures within EUS sub-regions are projected to continue to warm significantly (p-value<0.05). Under RCP8.5, SST of the Atlantic EUS is projected to increase by 0.81°C over 2010 2039 and 2.56°C over 2010 2099, Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 with similar increases projected for the Pacific EUS (Table SM30-4). Differences between RCPs for the two EUS become clear beyond mid-century, with warming of SST over 2010 2099 being 0.43°C and 0.46°C under RCP2.6 and 3.01°C and 3.03°C under RCP8.5, for Pacific and Atlantic EUS respectively (Table SM30-4). These projected increases in sea temperature will increase heat stress and ultimately irreversibly degrade marine ecosystems such as coral reefs (very likely). Further increases in atmospheric CO2 will cause additional decrease in pH and aragonite saturation of surface waters (adding to the low pH and aragonite saturation of upwelling conditions), with significant differences between emission trajectories by the middle of the century. These changes in ocean carbonate chemistry are very likely to negatively affecting some marine calcifiers, although many of the species from this region are adapted to the low aragonite and calcite saturation states that result from equatorial upwelling, albeit with much lower rates of calcification [Manzello, 2010; Friedrich et al., 2012]. A substantial risk exists with respect to the synergistic interactions between sea temperature and declining pH, especially as to how they influence a large number of key biological processes (Box CC-OA). There is low confidence in the current understanding of how (or if) climate change will influence the behavior of ENSO and other long-term climate patterns ([Collins et al., 2010]; WGI 12.4.4.2). There is also low agreement between different CMIP5 GCMs on how ocean warming will affect ENSO, with no significant change to ENSO amplitude in half the models examined, and both increasing and decreasing activity in others [Guilyardi et al., 2012]. These differences appear to be a consequence of the delicate balance within ENSO between dampening and amplifying feedbacks, and the different emphasis given to these processes within the different GCMs [Collins et al., 2010]. Other studies have looked at the interaction between the STG and EUS, and warming of surface waters in the Pacific, with at least one study projecting the possible expansion of the STG at the expense of the EUS [Polovina et al., 2011]. In the latter case, the area of equatorial upwelling within the North Pacific would decrease by 28%, and primary production and fish catch by 15%, by 2100. Many of the projected changes imply additional consequences for pelagic fisheries due to the migration of fish stocks deriving from changing distribution of particular sea temperatures [Lehodey et al., 2006; Lehodey et al., 2008; Cheung et al., 2010; Lehodey et al., 2011; Sumaila et al., 2011; Bell et al., 2013b]. These projections suggest that fisheries within EUS will experience increased vulnerability due to elevated variability in space and time as a result of climate change (low confidence). 30.5.3. Semi-Enclosed Seas Semi-Enclosed Seas (SES) represent a subset of ocean sub-regions that are largely land-locked and consequently heavily influenced by surrounding landscapes and climates [Healy and Harada, 1991]. In most cases, they support small but regionally significant fisheries (3.3% of global production; Table SM30-1, Figure 30-1b) and opportunities for other industries such as tourism. Five SES (all over 200,000 km with single entrances <120 km wide) are considered here. This particular geography results in reduced circulation and exchange with ocean waters, and jurisdictions for these water bodies are shared by two or more neighboring states. In many cases, the small volume and disconnected nature of SES (relative to coastal and oceanic environments) makes them highly vulnerable to both local and global stressors, especially with respect to the much reduced options for the migration of organisms as conditions change. 30.5.3.1. Observed Changes and Potential Impacts 30.5.3.1.1. Arabian Gulf The Arabian Gulf (also referred to as the Persian Gulf), along with Red Sea, is the world's warmest sea, with both extreme negative and positive temperature excursions (annual temperature range of 12 35°C). Like other SES, the Arabian Gulf is particularly vulnerable to changing environmental conditions as a result of its landlocked nature. Trends in SST were not significant over the period 1950 2009 (Table 30-1), which is probably due to long-term variability, and a consequence of regional and abrupt changes that occurred in the late 1980s [Conversi et al., 2010]. In keeping with this, recent (1985 2002) localized analyses (e.g., Kuwait Bay) show strong and significant warming trends (based in this case on AVHRR (NOAA) satellite data) of 0.6°C decade-1 [Al-Rashidi et al., 2009]. There is limited evidence and low agreement as to how this variability influences marine ecosystems and human activities Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 within the Arabian Gulf, although impacts on some ecosystem components (e.g., coral reefs) have been defined to some extent. The mass coral bleaching and mortality that occurred in 1996 and 1998 were a direct result of the sensitivity of reef-building corals to elevated sea temperatures (high confidence) ([Riegl, 2002; 2003]; Box CC-CR). These changes to coral reefs have resulted in a loss of fish species that feed on coral-associated invertebrates while herbivores and planktivorous fish abundances have increased (medium confidence) [Riegl, 2002]. Despite coral ecosystems in this sub-region being adapted to some of the highest temperatures in shallow seas on earth, anthropogenic climate change is driving higher frequencies and intensities of mass coral bleaching and mortality [Riegl et al., 2011]. Other biological changes (e.g., harmful algal blooms and fish kills, [Heil et al., 2001]) have been associated with the increasing sea temperatures of the Arabian Gulf, although attribution to increasing temperatures as opposed to other factors (e.g., water quality) is limited [Bauman et al., 2010]. 30.5.3.1.2. Red Sea Few studies have focused on attributing recent changes in the Red Sea ecosystems to climate change (including ocean acidification). The Red Sea warmed by 0.74°C from 1982 2006 [Belkin, 2009], although trends in the average SST, however, are not significant from 1950 2009 (p-value>0.05, Table 30-1) due to a high degree of variability when longer periods were examined (supplementary material in [Belkin, 2009]). The temperature of the warmest month of the year, however, showed a significant increase over the 60-year period (0.05°C decade-1; Table 30-1). Regional trends within the Red Sea may also differ, with at least one other study reporting higher rates of warming for the central Red Sea (1.46°C, relative to 1950 1997 NOAA Extended Reconstructed SST (ERSST) v3b climatology [Cantin et al., 2010]). Long-term monitoring of coral community structure and size over 20 years shows that average colony size of corals has declined (high confidence) and species latitudinal limits may have changed (medium confidence). The decline in average colony size is ascribed to heat-mediated bleaching as well as increases in coral diseases and Crown of Thorns Starfish (Acanthaster sp.) predation [Riegl et al., 2012]. The patterns of this decline correlate well with the pattern of recent heating in the Red Sea [Raitsos et al., 2011] with the biggest changes being seen in the southern part of the Red Sea. Skeletal growth of the long-lived massive coral Diploastrea heliopora has declined significantly, very likely due to warming temperatures (medium confidence) (p-value<0.05; [Cantin et al., 2010]. Cantin et al. [2010] proposed that the massive coral Diploastrea helipora will cease to grow in the central Red Sea by 2070 under SRES A1B and A2 (medium confidence), although this may not hold for other coral species. For example, an increase in linear extension of Porites corals, beginning in the 1980s, was recorded in the northern Red Sea [Heiss, 1996], where temperatures have increased by 0.74C from 1982-2006 [Belkin, 2009] suggesting that these corals were living in sub-optimal conditions (cooler waters). They may therefore benefit from elevated temperature before reaching their thermal threshold, at which point growth rates would be predicted to decline, as they are doing in other oceans. Riegl and Piller [2003] concluded that coral habitats at moderate depths in the Red Sea might provide important refugia from some aspects of climate change in the future (limited evidence). Silverman et al. [2007] quantified the sensitivity of net coral reef ecosystem calcification to changes in carbonate chemistry (pH, aragonite saturation). Their results demonstrate a strong negative effect of ocean acidification on ecosystem- scale calcification and decalcification, and show that small changes in carbonate dissolution could have large-scale implications for the long-term persistence of carbonate coral reef systems within the Red Sea [Silverman et al., 2007; Silverman et al., 2009]. 30.5.3.1.3. Black Sea The temperature of the surface waters of the Black Sea increased by 0.96C from 1982 2006 [Belkin, 2009], which is consistent with other studies (high confidence) [Buongiorno Nardelli et al., 2010; Bozkurt and Sen, 2011]. As with other SES (i.e., Arabian Gulf and Baltic, Mediterranean, and Red Seas), longer data sets do not reveal a significant trend due to large-scale variability prior to 1982, which may be due to the influence of AMO, NAO, and other long-term sources of variability (Table 30-1; supplementary material in Belkin, 2009). Buongiorno Nardelli et al. [2010] observed that short-term SST variability (week-month) is strongly influenced by interactions with the Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 overlying atmosphere, which itself is strongly influenced by the surrounding land temperatures. As with the Mediterranean and Red Seas, however, a significant upward trend in the temperature is recorded in the warmest month of the year over the period 1950 2009 (Table 30-1). Freshwater discharge from rivers draining into the Black Sea has remained more or less constant since the early 1960s [Ludwig et al., 2009]. Increasing water temperature has steadily eliminated the Cold Intermediate Layer (CIL; temperatures below 8°C) throughout the Black Sea basin over 1991 2003 (high confidence)[Oguz et al., 2003]. Reduced water column mixing and upwelling during warmer winter periods has reduced the supply of nutrients to the upper layers of the Black Sea [Oguz et al., 2003] and expanded areas of low O2 in the deeper parts of the Black Sea, which is the world's largest anoxic marine basin (high confidence) [Murray et al., 1989]. These changes coincided with the collapse of fish stocks and the invasion by the ctenophore Mnemiopsis leidyi in the 1980s [Oguz et al., 2008], while inputs of nutrients such as phosphate from the Danube River decreased strongly since 1992 1993 [Oguz and Velikova, 2010]. Environmental perturbations explain the declining levels of primary productivity, phytoplankton, bacterioplankton, and fish stocks in the Black Sea from the mid-1990s [Yunev et al., 2007; Oguz and Velikova, 2010]. The Black Sea system is very dynamic and is strongly affected by non-climate stressors in addition to climate change, making attribution of detected trends to climate change difficult. 30.5.3.1.4. Baltic Sea Temperatures in the highly dynamic Baltic Sea increased substantially since the early 1980s [Aleksandrov et al., 2009; Belkin, 2009], with increases of 1.35°C (1982 2006) being among the highest for any SES [Belkin, 2009]. Increases of this magnitude are not seen in longer records throughout the Baltic Sea (1861 2001, [MacKenzie et al., 2007; MacKenzie and Schiedek, 2007b; a]; 1900 1998, [Madsen and Hjerslev, 2009]). The salinity of the surface and near bottom waters of the Baltic Sea (e.g., Gdansk Basin, [Aleksandrov et al., 2009]; central Baltic [Fonselius and Valderrama, 2003; Möllmann et al., 2003] decreased from 1975 2000, due to changing rainfall and river run- off, and a reduction in the pulses of sea water (vital for oxygenation and related chemical changes) from the North Sea through its opening via the Kattegat (high confidence) [Samuelsson, 1996; Conley et al., 2009; Hänninen and Vuorinen, 2011]. There is a strong vertical zonation within the Baltic Sea in terms of the availability of O2. The shallow sub-regions of the Baltic are relatively well oxygenated. However, O2 levels are low in the deeper basins, producing conditions where organisms and ecosystems are exposed to prolonged hypoxia. The annual biomass of phytoplankton has declined almost threefold in the Baltic Transition Zone (Kattegat, Belt Sea) and Western Baltic Sea since 1979 [Henriksen, 2009], reputedly due to changing nitrogen loads in the Danish Straits (medium confidence) in addition to increasing sea temperature (very likely) [Madsen and Hjerslev, 2009]. Reduced phytoplankton production may have reduced the productivity of fisheries in the western Baltic Sea and the Transition Zone (low to medium confidence) [Chassot et al., 2007]. Decreasing salinity in the Baltic deep basins may also affect zooplankton reproduction, especially that of the copepod Pseudocalanus acuspes, contributing to density-dependent decrease in growth of the commercially important herring and sprat stocks (high confidence) [Möllmann et al., 2003; Möllmann et al., 2005; Casini et al., 2011]. The strong relationship between phytoplankton and fish production, and increasing sea temperature, decreasing salinity and other environmental factors, suggests that major changes in fisheries production will occur as sea temperatures increase and the hydrological cycle in the Baltic region changes (high confidence) [MacKenzie et al., 2012]. A combination of climate change-induced oceanographic changes (i.e., decreased salinity and increased temperatures), eutrophication, and overfishing have resulted in major changes in trophic structure in the deep basins of the Baltic Sea [Möllmann et al., 2009]. This had important implications for cod, a commercially important top-predator (medium confidence)[Lindegren et al., 2010]. 30.5.3.1.5. Mediterranean Sea The Mediterranean Sea is strongly linked to the climates of North Africa and Central Europe. SST within the Mediterranean increased by 0.43°C from 1957 2008 (supplementary material, [Belkin, 2009]), although analysis of data from 1950 2009 only detected a significant trend in summer temperature (0.11°C decade-1, p-value<0.05, Table 30-1) due to large fluctuations in SST prior to the 1980s. Surface temperatures increased in the Mediterranean Sea consistent with significant increases in SST at a number of monitoring sites (high agreement, robust evidence) (e.g. Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [Coma et al., 2009; Conversi et al., 2010; Calvo et al., 2011]). It is likely that temperatures, along with salinity, have also increased at depth (400 m or more) in the western Mediterranean Sea over the past 30-40 years which, when analyzed in the context of heat budget and water flux of the Mediterranean, is consistent with anthropogenic greenhouse warming [Bethoux et al., 1990; Rixen et al., 2005; Vargas-Yánez et al., 2010]. Large scale variability such as the AMO and NAO can obscure or accentuate the overall warming trend ([Marullo et al., 2011]; WGI 14.5.1, 14.7.6). Relatively warm episodes in the 1870s, 1930 1970s and since the mid-1990s, for example, exhibit an influence of the AMO [Kerr, 2000; Moron, 2003]. Reported temperature anomalies in the Mediterranean, often locally manifesting themselves as periods of low wind, increased water column stratification, and a deepening thermocline, are associated with positive phases of the NAO index [Molinero et al., 2005; Lejeusne et al., 2010]. Sea levels have increased rapidly in some areas over recent decades and are also strongly influenced by NAO phases. The rate has been approximately 3.4 mm yr-1 (1990 2009) in the North-west Mediterranean (high confidence) [Calvo et al., 2011]. These influences are reduced when measurements are pooled over longer time- scales, resulting in a lower rate of sea level rise [Massuti et al., 2008]. If the positive phase of the NAO is more frequent in the future ([Terray et al., 2004; Kuzmina et al., 2005]; WGI 14.4.2), then future sea level rise may be slightly suppressed due to atmospheric influences (medium confidence) [Jorda et al., 2012]. As temperatures have increased, the Mediterranean has become more saline (+0.035 0.040 psu from 1950 2000, [Rixen et al., 2005]) and the length of the thermal stratification period persisted twice as long in 2006 as it did in 1974 [Coma et al., 2009]. Conditions within the Mediterranean Sea changed abruptly and synchronously with similar changes across the North, Baltic, and Black Seas in the late 1980s [Conversi et al., 2010], which possibly explains the lack of trend in SES SST when examined from 1950 2009 (Table 30-1). These changes in physical conditions (increased temperature, higher sea level pressure, positive NAO index) also coincided with step-changes in the diversity and abundance of zooplankton, decreases in stock abundance of anchovies, decreases in the frequency of red tides , and increases in mucilage outbreaks [Conversi et al., 2010]. Mucilage outbreaks are strongly associated with warmer and more stratified water columns (high confidence), and lead to a greater abundance and diversity of marine microbes and potentially disease-causing organisms (likely) [Danovaro et al., 2009]. Increasing temperatures are also driving the northward spread of warm-water species (medium confidence) such as the sardine Sardinella aurita [Sabatés et al., 2006; Tsikliras, 2008], and have contributed to the fast spread of the invading Atlantic coral Oculina patagonia [Serrano et al., 2013]. The recent spread of warm-water species that have invaded through the Straits of Gibraltar and the Suez Canal into cooler northern areas is leading to the tropicalisation of Mediterranean fauna (high confidence) [Bianchi, 2007; Ben Rais Lasram and Mouillot, 2008; CIESM, 2008; Galil, 2008; 2011]. Warming since the end of the 1990s has accelerated the spread of tropical invasive species from the eastern Mediterranean basin ([Raitsos et al., 2010]; 23.6.5). In addition to general patterns of warming, periods of extreme temperatures have had large-scale and negative consequences for Mediterranean marine ecosystems. Unprecedented mass mortality events, that affected at least 25 prominent invertebrate species, occurred during the summers of 1999, 2003, and 2006 across hundreds of kilometers of coastline in the North-west Mediterranean Sea (very high confidence) [Cerrano et al., 2000; Garrabou et al., 2009; Calvo et al., 2011; Crisci et al., 2011]. Events coincided with either short periods (2 5 days: 2003, 2006) of high sea temperatures (27°C) or longer periods (30 40 days) of modestly high temperatures (24°C: 1999; [Bensoussan et al., 2010; Crisci et al., 2011]). Impacts on marine organisms have been reported in response to the extreme conditions during these events (e.g., gorgonian coral mortality [Coma et al., 2009], shoot mortality, and anomalous flowering of seagrasses (high confidence) [Diaz-Almela et al., 2007; Marba and Duarte, 2010]. The frequency and intensity of these types of heat stress events are expected to increase as sea temperatures increase (high confidence). Longer-term data series (over several decades) of changes in relative acidity of the Mediterranean Sea are scarce [Calvo et al., 2011; MerMex-Group, 2011]. Recent re-analysis, however, has concluded that the pH of Mediterranean waters has decreased by 0.05 0.14 pH units since the pre-industrial period (medium confidence) [Luchetta et al., 2010; Touratier and Goyet, 2011]. Anthropogenic CO2 has penetrated the entire Mediterranean water column, with the western basin being more contaminated than the eastern basin [Touratier and Goyet, 2011]. Studies that have explored the consequences of ocean acidification for the biology and ecology of the Mediterranean Sea are rare [Martin and Gatusso, 2009; Rodolfo-Metalpa et al., 2010; Movilla et al., 2012], although insights have Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 been gained by studying natural CO2 seeps at Mediterranean sites such as Ischia in Italy, where biodiversity decreases with decreasing pH towards the vents, with a notable decline in calcifiers. [Hall-Spencer et al., 2008]. Transplants of corals, mollusks, and bryozoans along the acidification gradients around seeps reveal a low level of vulnerability to CO2 levels expected over the next 100 years (low confidence) [Rodolfo-Metalpa et al., 2010 ; Rodolfo-Metalpa et al., 2011]. However, periods of high temperature can increase vulnerability to ocean acidification, thereby increasing the long-term risk posed to Mediterranean organisms and ecosystems as temperatures warm. Significantly, some organisms such as seagrasses and some macroalgae appeared to benefit from local ocean acidification [Hall-Spencer et al., 2008]. 30.5.3.2. Key Risks and Vulnerabilities SES are highly vulnerable to changes in global temperature on account of their small volume and landlocked nature. Consequently, SES will respond faster than most other parts of the Ocean (high confidence). Risks to ecosystems within SES are likely to increase as water columns become further stratified under increased warming, promoting hypoxia at depth and reducing nutrient supply to the upper water column (medium evidence, high agreement). The impact of rising temperatures on SES is exacerbated by their vulnerability to other human influences such as overexploitation, pollution, and enhanced run-off from modified coastlines. Due to a mixture of global and local human stressors, key fisheries have undergone fundamental changes in their abundance and distribution over the past 50 years (medium confidence). A major risk exists for SES from projected increases in the frequency of temperature extremes that drive mass mortality events, increasing water column stabilization leading to reduced mixing, and changes to the distribution and abundance of marine organisms. The vulnerability of marine ecosystems, fisheries, and human communities associated with the SES will continue to increase as global temperatures increase. Sea temperatures are very likely to increase in the five SES under moderate (RCP6.0) to high (RCP8.5) future scenarios. Under BAU (RCP8.5; Table SM30-3), sea temperatures in the SES are projected to increase by 0.93 1.24°C over 2010 2039 (Table SM30-4). Increases of 3.45 4.37°C are projected over 2010 2099, with the greatest increases projected for the surface waters of the Baltic Sea (4.37°C) and Arabian Gulf (4.26°C), and lower yet substantial amounts of warming in the Red Sea (3.45°C) (Table SM30-4). The heat content added to these small ocean regions is very likely to increase stratification, which will reduce the nutrient supply to the upper layers of the water column, reducing primary productivity and driving major changes to the structure and productivity of fisheries. Reduced mixing and ventilation, along with increased microbial metabolism, will very likely increase hypoxia and expand the number and extent of dead zones . Changing rainfall intensity (23.3, WGI 12.4.) can exert a strong influence on the physical and chemical conditions within SES, and in some cases will combine with other climatic changes to transform these areas. These changes are likely to increase the risk of reduced bottom-water O2 levels to Baltic and Black Sea ecosystems (due to reduced solubility, increased stratification, and microbial respiration), which is very likely to affect fisheries. These changes will increase the frequency and intensity of impacts arising from heat stress, based on responses to temperature extremes seen over the past 30 years, such as the mass mortality of benthic organisms that occurred in the Mediterranean Sea during the summers of 1999, 2003, and 2006, and the Arabian Gulf in 1996 and 1998. Extreme temperature events such as heat waves are projected to increase (high confidence) (23.2, [IPCC, 2012]). Similar projections to those outlined in 30.5.4.2 can be applied to the coral reefs of the Arabian Gulf and the Red Sea, where temperatures are very likely to increase above established thresholds for mass coral bleaching and mortality (very high confidence) (Figure 30-10). 30.5.4. Coastal Boundary Systems The Coastal Boundary Systems (CBS) are highly productive regions, comprising 10.6% of primary production and 28.0% of global fisheries production (Table SM30-1, Figure 30-1b). The CBS include the marginal seas of the North-west Pacific, Indian, and Atlantic Oceans, encompassing: the Bohai/Yellow Sea, East China Sea, South China Sea, and South-east Asian Seas (e.g., the Timor, Arafura, and Sulu Seas, and the northern coast of Australia) in the Pacific; the Arabian Sea, Somali Current system, East Africa coast, Mozambique Channel, and Madagascar in the Indian Ocean; and the Caribbean Sea and Gulf of Mexico in the Atlantic Ocean). Some CBS are dominated by Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 powerful currents such as the Kuroshio (Pacific), or are strongly influenced by monsoons (e.g., Asian-Australian and African monsoons). 30.5.4.1. Observed Changes and Potential Impacts Many ecosystems within the CBS are strongly affected by the local activities of often-dense coastal human populations. Activities such as the overexploitation of fisheries, unsustainable coastal development, and pollution have resulted in the wide-spread degradation of CBS ecosystems [Burke et al., 2002; Burke et al., 2011]. These influences have combined with steadily increasing ocean temperature and acidification to drive major changes to a range of important ecosystems over the past 50 years. Understanding the interactions between climate change and non-climate change drivers is a central part of the detection and attribution process within the CBS. Overall, the CBS warmed by 0.14 0.80°C from 1950 2009 (Table 30-1), although changes within the Gulf of Mexico/Caribbean Sea sub-region were not significant (p-value>0.05) over this period. Key sub-regions within the CBS such as the Coral Triangle and Western Indian Ocean warmed by 0.79 and 0.60°C, respectively, from 1950 2009 (Table 30-1). Rates of sea level rise vary from decreasing sea levels (-5 to -10 mm yr-1) to low (2 3 mm yr-1, Caribbean) to very high (10 mm yr-1, South-east Asia; Figure 30-5) rates of increase. Ocean acidification also varies from region to region (Figure SM30-2), and is influenced by oceanographic and coastal processes, which often have a large human component. 30.4.4.1.1. Bohai/Yellow Sea/East China Sea The Bohai Sea, Yellow Sea and the East China Sea (ECS) are shallow marginal seas along the edge of the North- west Pacific that are strongly influenced by the Kuroshio Current [Matsuno et al., 2009], the East Asian Monsoon (EAM), and major rivers such as the Yellow (Huang He) River and Yangtze (Changjiang) River. Upwelling of the Kuroshio sub-surface waters provides abundant nutrients that support high levels of primary productivity [Wong et al., 2000; Wong et al., 2001]. The ecosystems of the ECS are heavily affected by human activities (e.g., overfishing and pollution), which tend to compound the influence and consequences of climate change. SST within the ECS has increased rapidly since the early 1980s (high confidence) [Lin et al., 2005; Jung, 2008; Cai et al., 2011; Tian et al., 2012]. The largest increases in SST have occurred in the ECS in winter (1.96°C, 1955 2005) and in the Yellow Sea in summer (1.10°C, 1971 2006, [Cai et al., 2011]). These changes in SST are closely linked to a weakening of the EAM (e.g., [Cai et al., 2006; Tang et al., 2009; Cai et al., 2011]) and increasing warmth of the Kuroshio Current [Qi et al., 2010; Zhang et al., 2011; Wu et al., 2012]. At the same time, dissolved O2 has decreased [Lin et al., 2005; Jung, 2008; Qi et al., 2010], with an associated increase in the extent of the hypoxic areas in coastal areas of the Yellow Sea/ECS [Jung, 2008; Tang, 2009; Ning et al., 2011]. Primary productivity, biomass yields, and fish capture rates have experienced large changes within the ECS over the past decades (limited evidence, medium agreement, low confidence) [Tang et al., 2003; Lin et al., 2005; Tang, 2009]. Fluctuations in herring abundance appear to closely track SST shifts within the Yellow Sea [Tang, 2009]. For plankton and fish species, the proportions of warm-water species relative to warm-temperate species in the Changjiang River Estuary (extending to the southern Taiwan Strait) have changed in past decades [Zhang et al., 2005; Ma et al., 2009; Lin and Yang, 2011]. Northward shifts in catch distribution for some pelagic fish species in Korean waters were driven, in part, by warming SST (medium confidence, [Jung et al., 2013]. The frequency of harmful algal blooms (HAB) and blooms of the giant jellyfish Nemopilema nomurai in the offshore area of the ECS have increased and have been associated with ocean warming and other factors such as eutrophication [Ye and Huang, 2003; Tang, 2009; Cai and Tan, 2010]. While attribution of these changes to anthropogenic climate change is complicated by the increasing influence of non-climate related human activities, many of these changes are consistent with those expected as SST increases. Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.5.4.1.2. South China Sea The South China Sea (SCS) is surrounded by continental areas and a large number of islands, and is connected to the Pacific, ECS, and Sulu Sea by straits such as the Luzon and Taiwan Strait. The region is greatly influenced by cyclones/typhoons, and by the Pearl, Red, and Mekong Rivers. The region has a distinct seasonal circulation and is greatly influenced by the southwest monsoon (in summer), the Kuroshio Current and northeast monsoon (in winter). The SCS includes significant commercial fisheries areas and includes coral reefs, mangroves, and seagrasses. The surface waters of the SCS have been warming steadily from 1945 1999 with the annual mean SST in the central SCS increasing by 0.92°C (1950 2006, [Cai et al., 2009]), a rate similar to that observed for the entire Indo- Pacific/SE Asian CBS from 1950 2009 (0.80°C, Table 30-1). Significant freshening in the SCS intermediate layer since the 1960s has been observed [Liu et al., 2007]. The temperature change of the upper layers of the SCS has made a significant contribution to sea level variation, which is spatially non-homogeneous and varies in time [Li et al., 2002; Cheng and Qi, 2007; Liu et al., 2007]. Identifying the extent to which climate change is influencing the SCS is difficult due to confounding non-climate change factors and their interactions (e.g., local human pollution, over-exploitation together with natural climate variability such as EAM, ENSO, and PDO). Changing sea temperatures have influenced the abundance of phytoplankton, benthic biomass, cephalopod fisheries, and the size of demersal trawl catches in the northern SCS observed over the period 1976 2004 (limited evidence, medium agreement) [Ning et al., 2009]. Coral reefs and mangroves are degrading rapidly as a result of both climate change and non-climate change related factors (very likely) (Box CC-CR, [Chen et al., 2009; China-SNAP, 2011; Zhao et al., 2012]). Mass coral bleaching and mortality of coral reefs within the SCS were triggered by elevated temperatures in 1998 and 2007 [Yu et al., 2006; Li et al., 2011]. Conversely, warming enabled the establishment of a high latitude, non-carbonate, coral community in Daya Bay in the northern SCS, although this community has recently degraded due to increasing anthropogenic stresses [Chen et al., 2009; Qiu et al., 2010]. 30.5.4.1.3. South-east Asian Seas The South-east Asian Seas (SAS) include an archipelago of diverse islands that interact with the westward flow of the North Equatorial Current and the Indonesian Throughflow (Figure 30-1a). A large part of this region is referred to as the Coral Triangle [Veron et al., 2009]. The world's most biologically diverse marine area, it includes parts of Malaysia, Indonesia, the Philippines, Timor Leste, the Solomon Islands, and Papua New Guinea. SST increased significantly from 1985 2006 [Penaflor et al., 2009; McLeod et al., 2010], although with considerable spatial variation. Trends examined over longer periods (1950 2009) show significant warming (+0.80°C, p-value<0.05, Table 30-1). The sea level is rising by up to 10 mm yr-1 in much of this region [Church et al., 2004; Church et al., 2006; Green et al., 2010]. Like other tropical areas in the world, coral reefs within SAS have experienced periods of elevated temperature, which has driven several mass coral bleaching and mortality events since the early 1980s (high confidence) [Hoegh-Guldberg et al., 2009; McLeod et al., 2010] (Figure 30-10a). The most recent occurred during warm conditions in 2010 [Krishnan et al., 2011]. These changes are the result of increasing ocean temperatures and are very likely to be a consequence of anthropogenic climate change (high confidence) (Box CC- CR, WGI 10.4.1). Although calcification rates of some key organisms (e.g., reef-building corals; [Tanzil et al., 2009]) have slowed over the past two decades, it is not possible to conclude that the changes are due to ocean acidification. While a large part of the decline in coral reefs has been due to increasing local stresses (principally destructive fishing, declining water quality, and overexploitation of key reef species), projected increases in SST represent a major challenge for these valuable ecosystems (high agreement) [Burke et al., 2002; Burke and Maidens, 2004]. 30.5.4.1.4. Arabian Sea and Somali Current The Arabian Sea and Somali Current are relatively productive ocean areas, being strongly influenced by upwelling and the monsoonal system. Wind-generated upwelling enhances primary production in the western Arabian Sea Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [Prakash and Ramesh, 2007]. Several key fisheries within this region are under escalating pressure from both fishing and climate change. Sea surface temperature increased by 0.18°C and 0.26°C in the Arabian Sea and Somali Current, respectively, from 1982 2006 (HadSST2, [Rayner et al., 2003; Belkin, 2009]), which is consistent with the overall warming of the Western Indian Ocean portion of the CBS from 1950 2009 (0.60°C, Table 30-1). Salinity of surface waters in the Arabian Sea increased by 0.5 1.0% over the past 60 years (Figure 30-6c), due to increased evaporation from warming seas and contributions from the outflows of the saline Red Sea and Arabian Gulf. As in other tropical sub-regions, increasing sea temperatures have increased the frequency of mass coral bleaching and mortality within this region [Wilkinson and Hodgson, 1999; Goreau et al., 2000; Wilkinson, 2004]. The aragonite saturation horizon in both the Arabian Sea and Bay of Bengal is now 100 200 m shallower than in pre-industrial times as a result of ocean acidification (medium confidence) [Feely et al., 2004]. Shoaling of the aragonite saturation horizon is likely to affect a range of organisms and processes, such as the depth distribution of pteropods (zooplankton) in the western Arabian Sea (medium confidence) [Hitchcock et al., 2002; Mohan et al., 2006]. More than 50% of the area of oxygen minimum zones (OMZs) in the world s oceans occur in the Arabian Sea and Bay of Bengal and long-term measurements reveal that O2 concentrations are declining in this region (high confidence) [Helly and Levin, 2004; Karstensen et al., 2008; Stramma et al., 2010] (30.3.2.3). The information regarding the consequences of climate change within this region is undeveloped and suggests that important physical, chemical, and biological responses to climate change need to be the focus of further investigation. 30.5.4.1.5. East Africa coast and Madagascar The Western Indian Ocean strongly influences the coastal conditions associated with Kenya, Mozambique, Tanzania, Madagascar, La Réunion, Mayotte, and three archipelagos (Comoros, Mauritius, and the Seychelles). Sea temperatures in the Western Indian Ocean have increased by 0.60°C over 1950 2009 (high confidence) (p- value<0.05; Table 30-1), increasing the frequency of positive thermal anomalies which have triggered mass coral bleaching and mortality events across the region over the past two decades (high confidence,[Baker et al., 2008; Nakamura et al., 2011](CC-HS). Trends in changes in SST and surface salinity vary with location along the East African coastline, with faster rates at higher latitudes (Figure 30-2). Periods of heat stress over the past 20 years have triggered mass coral bleaching and mortality on coral reef ecosystems within this region [McClanahan et al., 2007; McClanahan et al., 2009a; McClanahan et al., 2009c; McClanahan et al., 2009b; Ateweberhan and McClanahan, 2010; Ateweberhan et al., 2011]. Steadily increasing sea temperatures have also produced anomalous growth rates in long-lived corals such as Porites (high confidence) [McClanahan et al., 2009b]. Differences in the susceptibility of reef-building corals to stress from rising sea temperatures has also resulted in changes to the composition of coral (high confidence) (p-value<0.05; [McClanahan et al., 2007]) and benthic fish communities (high confidence) (p-value<0.05; [Graham et al., 2008; Pratchett et al., 2011a]). These changes are very likely to alter species composition and potentially the productivity of coastal fisheries (robust evidence, high agreement, high confidence) [Jury et al., 2010], although there may be a significant lag between the loss of coral communities and the subsequent changes in the abundance and community structure of fish (p-value<0.05, [Graham et al., 2007]). Some of these potential changes can be adverted or reduced by interventions such as the establishment of marine protected areas and changes to fishing management [McClanahan et al., 2008; Cinner et al., 2009; Jury et al., 2010; MacNeil et al., 2010]. 30.5.4.1.6. Gulf of Mexico and Caribbean Sea The Gulf of Mexico and Caribbean Sea form a semi-contained maritime province within the Western Atlantic. These areas are dominated by a range of activities including mineral extraction, fishing, and tourism, which provide employment and opportunity for almost 75 million people who live in coastal areas of the US, Mexico, and a range of other Caribbean nations [Adams et al., 2004]. The Gulf of Mexico and Caribbean Sea have warmed by 0.31°C and 0.50°C and respectively from 1982 2006 (very likely) [Belkin, 2009]. Warming trends are not significant from 1950 2009 (Table 30-1), which may be partly due to spatial variability in warming patterns (30.5.3.1). The Caribbean region has experienced a sustained decrease in aragonite saturation state from 1996 2006 (very likely) Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [Gledhill et al., 2008]. Sea levels within the Gulf of Mexico and Caribbean Sea have increased at the rate of 2 3 mm y-1 from 1950 2000 [Church et al., 2004; Zervas, 2009]. Understanding influences of climate change on ocean ecosystems in this region is complicated by the confounding influence of growing human populations and activities. The recent expansion of the seasonal hypoxic zone, and the associated dead zone , in the Gulf of Mexico has been attributed to nitrogen inputs driven by land management [Turner and Rabalais, 1994; Donner et al., 2004] and changes to river flows, wind patterns, and thermal stratification of Gulf waters (high confidence) [Justiæ et al., 1996; Justiæ et al., 2007; Levin et al., 2009; Rabalais et al., 2009; Rabalais et al., 2010]. The increases in coastal pollution and fishing have potentially interacted with climate change to exacerbate impacts on marine ecosystems within this region (5.3.4, 29.3). These changes have often been abrupt and non-linear [Taylor et al., 2012]. A combination of local and global disturbances has driven a large-scale loss of reef-building corals across the Caribbean Sea since the late 1970s (high confidence) [Hughes, 1994; Gardner et al., 2003]. Record thermal stress in 2005 triggered the largest mass coral bleaching and mortality event on record for the region, damaging coral reefs across hundreds of km2 in the eastern Caribbean Sea (high confidence) [Donner et al., 2007; Eakin et al., 2010]. Although conditions in 2010 were milder than 2005, elevated temperatures still occurred in some parts of the Caribbean [Smith et al., 2013]. Increasing temperatures in the Caribbean have also been implicated in the spread of marine diseases [Harvell et al., 1999; Harvell et al., 2002; Harvell et al., 2004] and some introduced species (likely) [Firth et al., 2011]. As in other sub-regions, pelagic fish species are sensitive to changes in sea temperature and modify their distribution and abundance [Muhling et al., 2011]. Fish and invertebrate assemblages in the Gulf of Mexico have shifted deeper in response to SST warming over 1970s-2011 (medium confidence) [Pinsky et al., 2013]. Coral ecosystems in the Caribbean Sea are at risk from ocean acidification (very likely) [Albright et al., 2010; Albright and Langdon, 2011], although impacts are yet to be observed under field conditions. Ocean acidification may also be altering patterns of fish recruitment to coral reefs, although direct evidence for how this has affected Caribbean species is lacking (low confidence) [Dixson et al., 2008; Munday et al., 2009; Dixson et al., 2010]. 30.5.4.2. Key Risks and Vulnerabilities Worldwide, 850 million people live within 100 km of tropical coastal ecosystems such as coral reefs and mangroves deriving multiple benefits including food, coastal protection, cultural services, and income from industries such as fishing and tourism [Burke et al., 2011]. Marine ecosystems within the CBS are sensitive to increasing sea temperatures (Figure 30-10), although detection and attribution is complicated by the significant influence and interaction with non-climate change stressors (water quality, over-exploitation of fisheries, coastal degradation; Box CC-CR). Warming is likely to have changed the primary productivity of ocean waters, placing valuable ecosystems and fisheries within the ECS at risk (low to medium confidence). Other risks include the expansion of hypoxic conditions and associated dead zones in many parts of the CBS. Given the consequences for coastal ecosystems and fisheries, these changes are very likely to increase the vulnerability of coastal communities throughout the CBS. Sea temperatures are increasing within many parts of the CBS ecosystems (1950 2009, Table 30-1), and will continue to do so over the next few decades and century. Sea temperatures are projected to change by 0.34 0.50°C over the near-term (2010 2039) and by 0.23 0.74°C over the long-term (2010 2099) under the lowest RCP scenario (RCP2.6). Under BAU (RCP8.5), CBS sea temperatures are projected to increase by 0.62 0.85°C over tne near-term and 2.44 3.32°C over the long-term (Table SM30-4). Given the large-scale impacts (e.g., mass coral bleaching and mortality events) that have occurred in response to much smaller changes in the past over the CBS regions (0.14 0.80°C from 1950 2009, Table 30-1), the projected changes of 2.44 3.32°C over 2010-2099 are very likely to have large-scale and negative consequences for the structure and function of many CBS ecosystems (virtually certain), especially given the sensitivity of coral reefs to relatively small increases in temperature over the past three decades [Hoegh-Guldberg, 1999; Eakin et al., 2010; Lough, 2012]. Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 It is very likely that coral-dominated reef ecosystems within the CBS (and elsewhere) will continue to decline and will consequently provide significantly less ecosystem goods and services for coastal communities if sea temperatures increase by more than 1°C above current temperatures (Box CC-CR, Figure 30-10). Combining the known sensitivity of coral reefs within the Caribbean and Coral Triangle sub-regions [Strong et al., 1997; Hoegh- Guldberg, 1999; Strong et al., 2011], with the exposure to higher temperatures that are projected under medium (RCP4.5) to high (RCP8.5) scenarios, reveals that both coral reef-rich regions are virtually certain to experience levels of thermal stress (DHM>1) that cause coral bleaching every 1 2 years by the mid to late part of this century (robust evidence, high levels of agreement, very high confidence) (Figure 30-4b, c; Figure 30-10, Figure 30-12, Figure SM30-3; [van Hooidonk et al., 2013]). The frequency of mass mortality events (DHM>5, Figure 30-10a-c) also climbs towards events that occur every 1 2 years by the mid to late part of this century under low to high climate change scenarios (robust evidence, high agreement, very high confidence) [Hoegh-Guldberg, 1999; Donner et al., 2005; Frieler et al., 2012]. Mass mortality events that affect coral reefs will result in changes to community composition in the near-term (2010 2039) [Berumen and Pratchett, 2006; Adjeroud et al., 2009] and a continuing downward trend in reef-building coral stocks in the longer term [Gardner et al., 2003; Bruno and Selig, 2007; Baker et al., 2008]. It is virtually certain that composition of fisheries catches [Graham et al., 2007; Pratchett et al., 2011a] [Pratchett et al., 2008; Pratchett et al., 2011b] will change. The productivity of many fisheries will decrease (limited evidence, medium agreement) as waters warm, acidify, and stratify, and as crucial habitat such as coral reefs degrades (low confidence). These changes are very likely to increase the vulnerability of millions of people who live in coastal communities and depend directly on fisheries and other ecological goods and services [Hoegh-Guldberg et al., 2009; McLeod et al., 2010]. [INSERT FIGURE 30-10 HERE Figure 30-10: Annual maximum proportions of reef pixels with Degree Heating Months [Donner et al., 2007]; DHM) 1 (used for projecting coral bleaching; [Strong et al., 1997; Strong et al., 2011]) and DHM 5 (associated with bleaching across 100% of affected areas with significant mortality, [Eakin et al., 2010] for the period 1870 2009 for each of the six coral regions (Figure 30-4d) using the HadISST1.1 data set. The black line on each graph is the maximum annual area value for each decade over the period 1870 2009. This value is continued through 2010 2099 using CMIP5 data and splits into the four Representative Concentration Pathways (RCP2.6, 4.5, 6.0, and 8.5). DHM were produced for each of the four RCPs using the ensembles of CMIP models. From these global maps of DHM, the annual percentage of grid cells with DHM 1 and DHM 5 were calculated for each coral region. These data were then grouped into decades from which the maximum annual proportions were derived. The plotted lines for 2010 2099 are the average of these maximum proportion values for each RCP. Monthly SST anomalies were derived using a 1985 2000 maximum monthly mean (MMM) climatology derived in the calculations for Figure 30- 4. This was done separately for HadISST1.1, the CMIP5 models, and each of the four RCPs, at each grid cell for every region. DHMs were then derived by adding up the monthly anomalies using a 4-month rolling sum. Figure SM30-3 presents past and future sea temperatures for the six major coral reef provinces under historic, un-forced, RCP4.5 and RCP8.5 scenarios.] 30.5.5. Eastern Boundary Upwelling Ecosystems The Eastern Boundary Upwelling Ecosystems (EBUE) include the California, Peru/Humboldt, Canary/North-west Africa, and Benguela Currents. They are highly productive sub-regions with rates of primary productivity that may exceed 1000 g C m-2 yr-1. Although these provinces comprise less than 2% of the Ocean area, they contribute nearly 7% of marine primary production (Figure 30-1b) and more than 20% of the world s marine capture fisheries [Pauly and Christensen, 1995]. Catches in the EBUE are dominated by planktivorous sardine, anchovy, and horse/jack mackerel, and piscivorous benthic fish such as hake. Nutrient input from upwelling of cooler waters stimulates primary production that is transferred to mid and upper trophic levels, resulting in substantial fish, seabird, and marine mammal populations. As a result, the EBUE are considered hotspots of productivity and biodiversity [Block et al., 2011]. The high level of productivity is a result of large-scale atmospheric pressure gradients and wind systems that advect surface waters offshore leading to the upwelling of cold, nutrient-rich waters from depth (Box CC-UP) [Chavez and Messie, 2009; Chavez et al., 2011]. Upwelling waters are typically low in pH and high in CO2, Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and are likely to continue to enhance changes in pH and CO2 resulting from rising atmospheric CO2 [Feely et al., 2008; Gruber, 2011]. 30.5.5.1. Observed Changes and Potential Impacts There are extensive studies of the coupled climate-ecosystem dynamics of individual EBUE (e.g., California Current). Decadal variability poses challenges to the detection and attribution of changes within the EBUE to climate change, although there are a number of long-term studies that have been able to provide insight into the patterns of change and their causes. Like other ocean sub-regions, EBUE are projected to warm under climate change, with increased stratification and intensified winds as westerly winds shift poleward (likely). However, cooling has also been predicted for some EBUE, resulting from the intensification of wind-driven upwelling [Bakun, 1990]. The California and Canary Currents have warmed by 0.73 and 0.53°C (very likely) (p-value<0.05, 1950- 2009, Table 30-1), respectively, while no significant trend was detected in the sea surface temperatures of the Benguela (p-value = 0.44) and Humboldt Currents (p-value = 0.21) from 1950 2009 (Table 30-1). These trends match shorter-term trends for various EBUE using Pathfinder version 5 data [Demarcq, 2009]. These differences are likely to be the result of differences in the influence of long-term variability and the specific responses of coastal wind systems to warming, although an analysis of wind data over the same period did not pick up clear trends (low confidence, with respect to long-term wind trends) [Demarcq, 2009; Barton et al., 2013]. How climate change will influence ocean upwelling is central to resolving ecosystem and fishery responses within each EBUE. There is considerable debate, however, as to whether or not climate change will drive an intensification of upwelling (e.g., [Bakun et al., 2010; Narayan et al., 2010; Barton et al., 2013] in all regions. This debate is outlined in Box CC-UP. EBUE are also areas of naturally low pH and high CO2 concentrations due to upwelling, and consequently may be vulnerable to ocean acidification and its synergistic impacts [Barton et al., 2012]. A full understanding of the consequences of ocean acidification for marine organisms and ecosystems is discussed elsewhere (Box CC-OA, Box CC-UP, 6.2, 6.3.2, [Kroeker et al., 2013], WGI 6.4). 30.5.5.1.1. Canary Current Part of the North Atlantic STG, the Canary Current extends from northern Morocco southwestward to the North Atlantic Equatorial Current. It is linked with the Portugal Current (which is sometimes considered part of the Canary Current) upstream and extends downstream to the Atlantic Equatorial Current. The coastal upwelling system, however, is limited to a narrow belt along the Saharan west coast to the coast of Guinea, with the most intense upwelling occurring centrally, along the coasts of Mauritania (15 20 N) and Morocco (21 26° N). Total fish catches, comprising mainly coastal pelagic sardines, sardinellas, anchovies, and mackerel, have fluctuated around 2 million tons yr-1 since the 1970s (http://www.seaaroundus.org/lme/27.aspx). Contrasting with the other EBUE, fishing productivity is modest, probably due to the legacy of uncontrolled fishing in the 1960s [Arístegui et al., 2009]. Most observations suggest that the Canary Current has warmed since the early 1980s [Arístegui et al., 2009; Belkin, 2009; Demarcq, 2009; Barton et al., 2013], with analysis of HadISST1.1 data from 1950 2009 indicating warming of 0.53°C from 1950 2009 (p-value<0.05, Table 30-1). Gómez-Gesteira et al. [2008] suggest a 20% and 45% decrease in the strength of upwelling in winter and summer, respectively, from 1967 2006, consistent with a decrease in wind strength and direction over the past 60 years. More recently, [Barton et al., 2013] show no clear increasing or decreasing trend in wind strength over the past 60 years, and a lack of agreement among wind trends and variability from different wind products, (e.g. PFEL, ICOADS, WASWind). This study presents no evidence for changes in upwelling intensity, with the exception of upwelling off North-west Spain, where winds are becoming slightly less favorable. Alteration of wind direction and strength influences upwelling and hence nutrient concentrations, howevernutrient levels can also change in response to other variables such as the supply of iron- laden dust from the Sahara [Alonso-Pérez et al., 2011]. There is medium evidence and medium agreement that primary production in the Canary Current has decreased over the past two decades [Arístegui et al., 2009; Demarcq, 2009], in contrast to the nearby upwelling region off North-west Spain, where no significant trend was observed Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [Bode et al., 2011]. Satellite chlorophyll records (SeaWiFS, MODIS) are relatively short, making it difficult to distinguish the influence of warming oceans from longer-term patterns of variability [Arístegui et al., 2009; Henson et al., 2010]. Changing temperature has resulted in changes to important fisheries species. For example, Mauritanian waters have become more suitable as feeding and spawning areas for some fisheries species (e.g., Sardinella aurita) as temperatures increased [Zeeberg et al., 2008]. Clear attribution of these changes depends on the linkage between the Azores High and global temperature, and on longer records for both physical and biological systems as pointed out for data sets in general [Arístegui et al., 2009; Henson et al., 2010]. 30.5.5.1.2. Benguela Current The Benguela Current originates from the eastward-flowing, cold South Atlantic Current, flows northward along the southwest coast of Africa, and is bounded north and south by the warm-water Angola and Agulhas Currents, respectively. Upwelling is strongest and most persistent toward the center of the system in the Lüderitz-Orange River upwelling cell [Hutchings et al., 2009]. Fish catch reached a peak in the late 1970s of 2.8 million tons yr-1 (http://www.seaaroundus.org/lme/29/1.aspx), before declines in the northern Benguela, due to overfishing and interdecadal environmental variability, resulted in a reduced catch of around 1 million tons yr-1 (present) [Cury and Shannon, 2004; Heymans et al., 2004; Hutchings et al., 2009]. Offshore commercial fisheries currently comprise sardine, anchovy, horse mackerel, and hake, while the inshore artisanal and recreational fisheries comprise a variety of fish species mostly caught by hook and line. Most research on the Benguela Current has focused on fisheries and oceanography, with little emphasis on climate change. As with the other EBUE, strong interannual and interdecadal variability in physical oceanography make the detection and attribution of biophysical trends to climate change difficult. Nevertheless, the physical conditions of the Benguela Current are highly sensitive to climate variability over a range of scales, especially to atmospheric teleconnections that alter local wind stress [Hutchings et al., 2009; Leduc et al., 2010; Richter et al., 2010; Rouault et al., 2010]. Consequently, there is medium agreement, despite limited evidence [Demarcq, 2009], that upwelling intensity and associated variables (e.g., temperature, nutrient, and O2 concentrations) from the Benguela system will change as a result of climate change (Box CC-UP). The temperature of the surface waters of the Benguela Current did not increase from 1950 to 2009 (p-value>0.05, Table 30-1), although shorter records show an decrease in the south-central Benguela Current (0.35 0.55 C decade- 1 [Rouault et al., 2010] or an increase for the whole Benguela region (0.24C, Belkin [2009]). These differences between short versus long records indicate the substantial influence of long-term variability on the Benguela system [Belkin, 2009]. Information on other potential consequences of climate-change within the Benguela system is sparse. Sea-level rise is similar to the global mean, although it has not been measured rigorously within the Benguela [Brundrit, 1995; Veitch, 2007]. Although upwelling water in the northern and southern portions of the Benguela Current exhibits elevated and suppressed pCO2, respectively [Santana-Casiano et al., 2009]), the consequences of changing upwelling intensity remain poorly explored with respect to ocean acidification. Finally, while periodic hypoxic events in the Benguela system are largely driven by natural advective processes, these may be exacerbated by future climate change [Monteiro et al., 2008; Bakun et al., 2010]. Despite its apparent sensitivity to environmental variability, there is limited evidence of ecological changes in the Benguela Current EBUE due to climate change [Poloczanska et al., 2013]. For example, pelagic fish [Roy et al., 2007], benthic crustaceans [Cockcroft et al., 2008], and seabirds [Crawford et al., 2008] have demonstrated general eastward range shifts around the Cape of Good Hope. Although these may be associated with increased upwelling along the South African south coast, specific studies that attribute these changes to anthropogenic climate change are lacking. Trawl surveys of demersal fish and cephalopod species showed consistently predictable hotspots of species richness over a 20 30 year study period (the earliest surveys since 1984 off South Africa) that were associated with greater depths and cooler bottom waters [Kirkman et al., 2013]. However, major changes in the structure and function of the demersal community have been shown in some parts of the Benguela in response to environmental change e.g., due to predominately fishing pressure in the 1960s and environmental forcing in the early 2000s in the southern Benguela [Howard et al., 2007], therefore changes driven by climate change may eventually affect the persistence of these biodiversity hotspots [Kirkman et al., 2013]. Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.5.5.1.3. California Current The California Current spans ~23° of latitude from central Baja California, Mexico, to central British Columbia, Canada, linking the North Pacific Current (Westwind Drift) with the North Equatorial and Kuroshio Currents to form the North Pacific Gyre. High productivity driven by advective transport and upwelling [Hickey, 1979; Chelton et al., 1982; Checkley and Barth, 2009; Auad et al., 2011] supports well-studied ecosystems and fisheries. Fish catches from the California Current have been approximately 0.6 million tonnes yr-1 since 1950 (http://www.seaaroundus.org/lme/3.aspx), which makes it the lowest catch of the four EBUE. The ecosystem supports the foraging and reproductive activities of 2 6 million seabirds from around 100 species [Tyler et al., 1993]. Marine mammals are diverse and relatively abundant, including recovering populations of humpback whales, among others [Barlow et al., 2008]. The average temperature of the California Current warmed by 0.73°C from 1950 2009 (p-value <0.05, Table 30-1) and by 0.14 0.80°C from 1985 2007 [Demarcq, 2009]. Like other EBUE, the California Current is characterized by large-scale interannual and interdecadal climate-ecosystem variability [McGowan et al., 1998; Hare and Mantua, 2000; Chavez et al., 2003; Checkley and Barth, 2009]. During an El Nino, coastally-trapped Kelvin waves from the tropics deepen the thermocline, thereby severely reducing upwelling and increasing ocean temperatures from California to Washington [Peterson and Schwing, 2003; King et al., 2011]. Atmospheric teleconnections to the tropical Pacific alter wind stress and coastal upwelling. Therefore, the ENSO is intimately linked with Bakun s (1990) upwelling intensification hypothesis (Box CC-UP). Interdecadal variability in the California Current stems from variability in the Pacific-North America pattern [Overland et al., 2010], which is influenced by the PDO [Mantua et al., 1997; Peterson and Schwing, 2003] and the NPGO [Di Lorenzo et al., 2008]. The major effects of the PDO and NPGO appear north of 39°N [Di Lorenzo et al., 2008; Menge et al., 2009]. There is robust evidence and medium agreement that the California Current has experienced a decrease in the number of upwelling events (23 40%), but an increase duration of individual events resulting in an increase of the overall magnitude of upwelling events from 1967 2010 (high confidence) [Demarcq, 2009; Iles et al., 2012]. This is consistent with changes expected under climate change yet remains complicated by the influence of decadal-scale variability (low confidence) [Iles et al., 2012]. Oxygen concentrations have also undergone large and consistent decreases from 1984 2006 throughout the California Current, with the largest relative decreases occurring below the thermocline (21% at 300 m). The hypoxic boundary layer (<60 umol kg-1) has also shoaled by up to 90 m in some regions [Bograd et al., 2008]. These changes are consistent with the increased input of organic carbon into deeper layers from enhanced upwelling and productivity, which stimulates microbial activity and results in the drawdown of O2 (likely), [Bakun et al., 2010] but see also [McClatchie et al., 2010; Koslow et al., 2011]; WGI 3.8.3). These changes are likely to have reduced the available habitat for key benthic communities as well as fish and other mobile species [Stramma et al., 2010]. Increasing microbial activity will also increase the partial pressure of CO2, decreasing pH and the carbonate chemistry of seawater. Together with the shoaling of the saturation horizon, these changes have increased the incidence of low O2 and low pH water flowing onto the continental shelf (high confidence) (40 120 m, [Feely et al., 2008]), causing problems for industries such as the shellfish aquaculture industry [Barton et al., 2012]. 30.5.5.1.4. Humboldt Current The Humboldt Current is the largest of the four EBUE, covering an area larger than the other three combined. It comprises the eastern edge of South Pacific Gyre, linking the northern part of the Antarctic Circumpolar Current with the Pacific South Equatorial Current. Although the primary productivity per unit area is modest compared to that of the other EBUE, the Humboldt Current system has very high levels of fish production. Current catches are in line with a long-term average (since the 1960s) of 8 million tons yr-1 (http://www.seaaroundus.org/lme/13/1.aspx), although decadal-scale variations range from 2.5 13 million tons yr-1. While the anchovies currently contribute 80% of the total catch, they alternate with sardines on a multi-decadal scale, with their dynamics mediated by the approach and retreat of sub-tropical waters to and from the coast [Alheit and Bakun, 2010]. This variability does not Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 appear to be altering due to anthropogenic climate change. Thus, from the late 1970s to the early 1990s, sardines were more important [Chavez et al., 2003]. The other major commercial fish species are jack mackerel among the pelagic fish, and hake among the demersal fish. The Humboldt Current EBUE did not show an overall warming trend in SST over the last 60 years (p-value>0.05, Table 30-1), which is consistent with other data sets (1982 2006, HadISST1.1, [Belkin, 2009]; 1985-2007, Pathfinder, [Demarcq, 2009]). Wind speed has increased in the central portions of the Humboldt Current, although wind has decreased in its southern and northern sections [Demarcq, 2009]. The lack of a consistent warming signal may be due to the strong influence of adjacent ENSO activity exerting opposing drivers on upwelling and which, if they intensify, would decrease temperatures (limited evidence, medium agreement). Similar to the Canary Current EBUE, however, there was a significant increase in the temperatures of the warmest month of the year over the period 1950 2009 (p-value<0.05, Table 30-1). Primary production is suppressed during warm El Nino events and amplified during cooler La Nina phases, these changes then propagate through to higher trophic levels [Chavez et al., 2003; Tam et al., 2008; Taylor et al., 2008]. However, in addition to trophic changes, there is also a direct thermal impact on organisms, which varies depending on the thermal adaptation window for each species (high confidence). A 37-year zooplankton time series for the coast of Peru showed no persistent trend in abundance and diversity [Ayón et al., 2004], although observed shifts coincided with the shifts in the regional SST. As for other EBUE, there is lack of studies that have rigorously attempted to detect and attribute changes to anthropogenic climate change, although at least two studies [Mendelssohn and Schwing, 2002; Gutierrez et al., 2011] provide additional evidence that the northern Humboldt Current has cooled (due to upwelling intensification) since the 1950s, a trend matched by increasing primary production. This is not entirely consistent with the lack of significant change over the period 1950 2009 (p- value>0.05, Table 30-1). Nevertheless, these relationships are likely to be complex in their origin, especially in their sensitivity to the long-term changes associated with ENSO and PDO, and the fact that areas within the Humboldt Current EBUE may be showing different behaviors. 30.5.5.2. Key Risks and Vulnerabilities EBUE are vulnerable to changes that influence the intensity of currents, upwelling, and mixing (and hence changes in SST, wind strength and direction), as well as O2 content, carbonate chemistry, nutrient content, and the supply of organic carbon to deep offshore locations (robust evidence, high agreement, high confidence). The extent to which any particular EBUE is vulnerable to these factors depends on location (Figure 3 from Gruber [2011] and other factors such as alternative sources of nutrient input and fishing pressure [Bakun et al., 2010]. This complex interplay between regional and global drivers means that our understanding of how factors such as upwelling within the EBUE will to respond to further climate change is uncertain (Box CC-UP, [Rykaczewski and Dunne, 2010]). In the GCM ensembles examined (Table SM30-3), modest rates of warming (0.22 0.93°C) occur within the four EBUEs in the near-term. Over 2010-2099, however, EBUE SSTs warm by 0.07 1.02C under RCP2.6, and 2.52 3.51C under RCP8.5 (Table SM30-4). These high temperatures have the potential to increase stratification of the water column and substantially reduce overall mixing in some areas. In contrast, the potential strengthening of coastal wind systems would intensify upwelling and stimulate primary productivity through the increased injection of nutrients into the photic zone of the EBUE (Box CC-UP). Garreaud and Falvey [2009] explored how wind stress along the South American coast would change by 2100 under B2 and A2 IPCC scenarios. Using an ensemble of 15 GCMs, southerly wind systems favoring upwelling increased along the sub-tropical coast of South America, extending and strengthening conditions for upwelling. Changes in the intensity of upwelling within the EBUE will drive fundamental changes to the abundance, distribution, and viability of resident organisms, although an understanding of their nature and direction is limited. In some cases, large-scale decreases in primary productivity and dependent fisheries are projected to occur for EBUE ecosystems [Blanchard et al., 2012], while other projections question the strong connection between primary productivity and fisheries production [Arístegui et al., 2009]. Increased upwelling intensity also has potential disadvantages. Elevated primary productivity may lead to decreasing trophic transfer efficiency, thus increasing the Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 amount of organic carbon exported to the seabed, where it is virtually certain to increase microbial respiration and hence increase O2 stress [Weeks et al., 2002; Bakun et al., 2010]. Increased wind stress may also increase turbulence, breaking up food concentrations (affecting trophic transfer), or causing excessive offshore advection, which could remove plankton from shelf habitats. The central issue for the EBUE is therefore whether or not upwelling will intensify and, if so, whether the negative consequences (e.g., reduced O2 and elevated CO2) associated with upwelling intensification will outweigh potential benefits from increased primary production and fisheries catch. 30.5.6. Sub-Tropical Gyres Sub-Tropical Gyres (STG) dominate the Pacific, Atlantic, and Indian Oceans (Figure 30-1a), and consist of large stable water masses that circulate clockwise (northern hemisphere) and anticlockwise (southern hemisphere) due to the Coriolis Effect. The oligotrophic areas at the core of the STG represent one of the largest habitats on Earth, contributing 21.2% of ocean primary productivity and 8.3% of the global fish catch (Figure 30-1b, Table SM30-1). A number of small island nations are found within this region. While many of the observed changes within these nations have been described in previous chapters (e.g., 5.3-4, 29.3 5), region-wide issues and consequences are discussed here due to the strong linkages between ocean and coastal issues. 30.5.6.1. Observed Changes and Potential Impacts The central portions of the STG are oligotrophic (Figure SM30-1. Temperatures within the STG of the North Pacific (NPAC), South Pacific (SPAC), Indian Ocean (IOCE), North Atlantic (NATL), and South Atlantic (SATL) have increased at rates of 0.020, 0.024, 0.032, 0.025, and 0.027°C yr-1 from 1998 2010, respectively ([Signorini and McClain, 2012]. This is consistent with increases observed from 1950 2009 (0.25 0.67C, Table 30-1). However differences among studies done over differing time-periods emphasize the importance of long-term patterns of variability. Salinity has decreased across the North and South Pacific STG (Figure 30-6c, WGI 3.3.3.1), consistent with warmer sea temperatures and an intensification of the hydrological cycle [Boyer, 2005]. The North and South Pacific STGs have expanded since 1993 (high confidence), with these changes likely being the consequence of a combination of wind forcing and long-term variability ([Parrish et al., 2000]; (WGI 3.6.3). Chlorophyll levels, as determined by remote-sensing of ocean color (Box CC-UP), have decreased in the NPAC, IOCE, and NATL by 9%, 12%, and 11%, respectively (p-value<0.5; [Signorini and McClain, 2012]) over and above the inherent seasonal and interannual variability from 1998 2010 [Vantrepotte and Mélin, 2011]. Chlorophyll levels did not change in the remaining two gyres (SPAC and SATL, and confirmed for SPAC by [Lee and McPhaden, 2010; Lee et al., 2010]). Furthermore, over the period 1998 2007, median cell diameter of key phytoplankton species exhibited statistically significant linear declines of about 2% in the North and South Pacific, and 4% in the North Atlantic Ocean [Polovina and Woodworth, 2012]. Changes in chlorophyll and primary productivity in these sub-regions have been noted before [McClain et al., 2004; Gregg et al., 2005; Polovina et al., 2008] and are influenced by seasonal and longer-term sources of variability (e.g., ENSO, PDO, 6.3.4, Figure 6.9). These changes represent a significant expansion of the world s most unproductive waters, although caution must be exercised given the limitations of satellite detection methods (Box CC-PP) and the shortness of records relative to longer-term patterns of climate variability. There is high confidence that changes that reduce the vertical transport of nutrients into the euphotic zone (e.g., decreased wind speed, increasing surface temperatures, and stratification) will reduce the rate of primary productivity and hence fisheries. 30.5.6.1.1. Pacific Ocean STG Pacific climate is heavily influenced by the position of the Intertropical Convergence Zone (ITCZ) and the South Pacific Convergence Zone (SPCZ), which are part of the ascending branch of the Hadley circulation (WGI 14.3.1). These features are also strongly influenced by interannual to interdecadal climate patterns of variability including ENSO and PDO. The current understanding of how ENSO and PDO will change as average global temperatures increase is not clear (low confidence) ([Collins et al., 2010], WGI 12.4.4.2). The position of both the ITCZ and Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 SPCZ vary seasonally and with ENSO [Lough et al., 2011], with a northward migration during the northern hemisphere summer and a southward migration during the southern hemisphere summer. These changes, along with the West Pacific Monsoon, determine the timing and extent of the wet and dry seasons in SPAC and NPAC sub- regions [Ganachaud et al., 2011]. Tropical cyclones are prominent in the Pacific (particularly the western Pacific), and CBS sub-regions between 10° 30° north and south of the equator, although the associated storm systems may occasionally reach higher latitudes. Spatial patterns of cyclones vary with ENSO, spreading out from the Coral Sea to the Marquesas Islands during El Nino and contracting back to the Coral Sea, New Caledonia, and Vanuatu during La Nina [Lough et al., 2011]. Historically, there have been almost twice as many land-falling tropical cyclones in La Nina as opposed to El Nino years off the east coast of Australia, with a declining trend in the number of severe tropical cyclones from 0.45 per year in the early 1870s to 0.17 per year in recent times [Callaghan and Power, 2011]. The Pacific Ocean underwent an abrupt shift to warmer sea temperatures in the mid-1970s as a result of both natural (e.g., Interdecadal Pacific Oscillation (IPO) and climate forcing (high confidence)[Meehl et al., 2009]. This change coincided with changes to total rainfall, rain days, and dry spells across the Pacific, with the direction of change depending on the location relative to the SPCZ. Countries such as the Cook Islands, Tonga, Samoa and American Samoa, and Fiji tend to experience drought conditions as the SPCZ (with cooler sea temperatures) moves toward the northeast during El Nino (high confidence). The opposite is true during La Nina conditions. The consequences of changing rainfall on the countries of the Pacific STG are discussed in greater detail elsewhere (5.4, 29.3, Table 29.1). While these changes are due to different phases of long-term variability in the Pacific, they illustrate the ramifications and sensitivity of the Pacific to changes in climate change. Elevated sea temperatures within the Pacific Ocean have increased the frequency of widespread mass coral bleaching and mortality since the early 1980s (very high confidence, [Hoegh-Guldberg and Salvat, 1995; Hoegh- Guldberg, 1999; Mumby et al., 2001; Baker et al., 2008; Donner et al., 2010]. There are few, if any, scientific records of mass coral bleaching and mortality prior to this period (high confidence [Hoegh-Guldberg, 1999]. Rates of decline in coral cover on coastal coral reef ecosystems range between 0.5 2.0% per year depending on the location within the Indo-Pacific region (high confidence, [Bruno and Selig, 2007; Hughes et al., 2011; Sweatman et al., 2011; De ath et al., 2012]. The reasons for this decline are complex and involve non-climate change related factors (e.g., coastal pollution and overfishing) as well as global warming and possibly acidification. A recent comprehensive analysis of the ecological consequences of coral bleaching and mortality concluded that bleaching episodes have resulted in catastrophic loss of coral reefs in some locations, and have changed coral community structure in many others, with a potentially critical influence on the maintenance of biodiversity in the marine tropics (high confidence, [Baker et al., 2008]. Increasing sea levels have also caused changes in seagrass and mangrove systems. Gilman et al. [2007] found a reduction in mangrove area with sea level rise, with the observed mean landward recession of three mangrove areas over four decades being 25, 64, and 72 mm yr-1, 12 37 times faster than the observed rate of sea level rise. Significant interactions exist between climate change and coastal development, where migration shoreward depends on the extent to which coastlines have been modified or barriers to successful migration have been established. Changes in sea temperature also lead to changes in the distribution of key pelagic fisheries such as skipjack tuna (Katsuwonus pelamis), yellowfin tuna (Thunnus albacares), big-eye tuna (T. obesus) and South Pacific albacore tuna (T. alalunga), which make up the majority of key fisheries in the Pacific Ocean. Changes in distribution and recruitment in response to changes in sea temperature as result of ENSO demonstrate the close association of pelagic fish stocks and water temperature. The shift in habitat for top predators in the northeast Pacific was examined by Hazen et al. [2012], who used tracking data from 23 marine species and associated environmental variables to predict changes of up to 35% in core habitat for these species within the north Pacific. Potential habitats are predicted to contract for the blue whale, salmon shark, loggerhead turtle, and blue and mako sharks, while potential habitats for the sooty shearwater, black -footed albatross, leatherback turtle, white shark, elephant seal, and albacore, bluefin and yellowfin tuna are predicted to expand [Hazen et al., 2012]. However, expansion of OMZs in the Pacific STG is predicted to compress habitat (depth) for hypoxia-intolerant species such as tuna [Stramma et al., 2010; Stramma et al., 2012]. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Reduction of ocean productivity of the STG [Sarmiento et al., 2004; Signorini and McClain, 2012] reduces the flow of energy to higher trophic levels such as those of pelagic fish [Le Borgne et al., 2011]. The distribution and abundance of fisheries stocks such as tuna are also sensitive to changes in sea temperature, and hence long-term variability such as ENSO and PDO. The redistribution of tuna in the western central equatorial region has been related to the position of the oceanic convergence zones, where the warm pool meets the cold tongue of the Pacific. These changes have been reliably reproduced by population models that use temperature as a driver of the distribution and abundance of tuna [Lehodey et al., 1997; Lehodey et al., 2006]. Projections of big-eye tuna (T. obesus) distributions under SRES A2 show an improvement in spawning and feeding habitats to 2100 in the eastern tropical Pacific and declines in the western tropical Pacific leading to an eastern displacement of tuna stocks [Lehodey et al., 2008; Lehodey et al., 2010b]. 30.5.6.1.2. Indian Ocean STG Like the Pacific Ocean, the Indian Ocean plays a crucial role in global weather patterns, with teleconnections throughout Africa, Australasia, Asia, and the Americas (e.g., [Clark et al., 2000; Manhique et al., 2011; Meehl and Arblaster, 2011; Nakamura et al., 2011]. Increasing sea level, temperature, storm distribution and intensity, and changing carbonate chemistry all influence the broad range of physical, chemical, and biological aspects of the Indian Ocean. Coral reef ecosystems in the Indian Ocean gyre system were heavily affected by record positive sea temperature anomalies seen in the southern hemisphere between February April 1998 (robust evidence, high agreement, high confidence) [Ateweberhan et al., 2011]. Coral cover across the Western Indian Ocean declined by an average of 37.7% after the 1998 heat stress event [Ateweberhan et al., 2011]. Responses to the anomalously hot conditions in 1998 varied between sub-regions, with the central Indian Ocean islands (Maldives, Seychelles, Chagos, and Lakshadweep) experiencing major decreases in coral cover directly after the 1998 event (from 40 53% coral cover in 1977 1997 to 7% in 1999 2000) (high confidence) [Ateweberhan et al., 2011]. Coral reefs lining the islands of southern India and Sri Lanka experienced similar decreases in coral cover (45%, 1977 1997 to 12%, 1999 2000). Corals in the southwestern Indian Ocean (Comoros, Madagascar, Mauritius, Mayotte, Réunion and Rodrigues) showed less impact (44%, 1977 1997 to 40%, 1999 2000). Recovery from these increases in mortality has been variable, with sites such as those around the central Indian Ocean islands exhibiting fairly slow recovery (13% by 2001 2005) while those around southern India and Sri Lanka are showing much higher rates (achieving a mean coral cover of 37% by 2001 2005, [Ateweberhan et al., 2011]). These changes to the population size of key reef-building species will drive major changes in the abundance and composition of fish populations in coastal areas, and affect other ecosystem services that are important for underpinning tourism and coastal protection (medium confidence, Box CC-CR). Fisheries that exploit tuna and other large pelagic species are very valuable to many small island states within the Indian Ocean). As with Pacific fisheries, the distribution and abundance of large pelagic fish in the Indian Ocean is greatly influenced by sea temperature. The anomalously high sea temperatures of 1997 98 (leading to a deepening of the mixed layer in the west and a shoaling in the east) coincided with anomalously low primary production in the western Indian Ocean and a major shift in tuna stocks (high confidence) ([Menard et al., 2007; Robinson et al., 2010]. Fishing grounds in the Western Indian Ocean were deserted and fishing fleets underwent a massive shift toward the eastern basin, which was unprecedented for the tuna fishery (high confidence). As a result of these changes, many countries throughout the Indian Ocean lost significant tuna-related revenue [Robinson et al., 2010]. In 2007, tuna fishing revenue was again reduced by strong surface warming and deepening of the mixed layer, and associated with a modest reduction in primary productivity in the west. These trends highlight the overall vulnerability of tuna fishing countries in the Indian Ocean to climate variability, a situation similar to the other major oceans of the world. 30.5.6.1.3. Atlantic Ocean STG SST has increased within the two STG of the Atlantic Ocean over the last two decades [Belkin, 2009; Signorini and McClain, 2012]. Over longer periods of time (1950 2009), trends in average temperature are not significant for the North Atlantic STG (p-value>0.05) while they remain so for the South Atlantic STG (very likely) (0.08°C decade-1, Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 p-value<0.05, Table 30-1). In both cases, however, temperatures in the coolest and warmest months increased significantly (Table 30-1). The difference between these studies (i.e., over 10 30 years versus 60 years) emphasizes the importance of long-term patterns of variability in the North Atlantic region. Variability in SST at a period of about 60 80 years is associated with the Atlantic Multi-decadal Oscillation (AMO) [Trenberth and Shea, 2006]. Sea surface temperatures influence hurricane activity (very likely) with recent record SST associated with record hurricane activity in 2005 in the Atlantic [Trenberth and Shea, 2006] and mass coral bleaching and mortality in the eastern Caribbean (high confidence, [Eakin et al., 2010]. In the former case, analysis concluded that 0.1°C of the SST anomaly was attributable to the state of the AMO while 0.45°C was due to ocean warming as a result of anthropogenic influences [Trenberth and Shea, 2006]. These changes have influenced the distribution of key fishery species as well the ecology of coral reefs in Bermuda [Wilkinson and Hodgson, 1999; Baker et al., 2008] and in the eastern Caribbean [Eakin et al., 2010]. Small island nations such as Bermuda depend on coral reefs for fisheries and tourism and are vulnerable to further increases in sea temperature that cause mass coral bleaching and mortality (high confidence, Box CC-CR, Figure 30-10). As with the other STG, phytoplankton communities and pelagic fish stocks are sensitive to temperature changes that have occurred over the past several decades. Observation of these changes has enabled development of models that have a high degree of accuracy in projecting the distribution and abundance of these elements within the Atlantic region in general [Cheung et al., 2011]. 30.5.6.2. Key Risks and Vulnerabilities Sea surface temperatures of the vast STG of the Atlantic, Pacific, and Indian Oceans are increasing, which is very likely to increase stratification of the water column. In turn, this is likely to reduce surface concentrations of nutrients and, consequently, primary productivity (medium confidence) (Box CC-PP). Warming is projected to continue (Table SM30-4), with substantial increases in the vulnerability and risk associated with systems that have been observed to change so far (high confidence) (Figure 30-12). Under RCP2.6, the temperatures of the STG are projected to increase by 0.17 0.56°C in the near-term (over 2010 2039) and between -0.03 0.90°C in the long-term (over 2010-2099) (Table SM30-4). Under RCP8.5, however, surface temperatures of the world s STG are projected to be 0.45 0.91°C warmer in the near-term and 1.90 3.44°C warmer in the long-term (Table SM30-4). These changes in temperature are very likely to increase water column stability, reduce the depth of the mixed layer, and influence key parameters such as nutrient availability and O2 concentrations. It is not clear as to how longer-term sources of variability such as ENSO and PDO will change (WGI 14.4, 14.7.6) and ultimately influence these trends. The world's most oligotrophic ocean sub-regions are likely to continue to expand over coming decades, with consequences for ecosystem services such as gas exchange, fisheries, and carbon sequestration. Polovina et al. [2011] explored this question for the North Pacific using a climate model that included a coupled ocean biogeochemical component to investigate potential changes under an SRES A2 scenario (~RCP6.0 8.5; Figure 1.5 [Rogelj et al., 2012]). Model projections indicated the STG expanding by approximately 30% by 2100, driven by the northward drift of the mid-latitude westerlies and enhanced stratification of the water column. The expansion of the STG occurred at the expense of the equatorial upwelling and other regions within the North Pacific. In the North Pacific STG, the total primary production is projected to decrease by 10-20% and large fish catch by 19 29% by 2100 under SRES A2 [Howell et al., 2013; Woodworth-Jefcoats et al., 2013]. However, our understanding of how large-scale eddy systems will change in a warming world is incomplete, as are the implications for primary productivity of these large and important systems (Box CC-PP, Box CC-UP). Understanding how storm frequency and intensity will change represents a key question for many countries and territories within the various STG. Projections of increasing sea temperature are likely to change the behavior of tropical cyclones. At the same time, the maximum wind speed and rainfall associated with cyclones is likely to increase, although future trends in cyclones and severe storms are very likely to vary from region to region (WGI 14.6). Patterns such as temporal clustering can have a strong influence on the impact of tropical cyclones on ecosystems such as coral reefs [Mumby et al., 2011], although how these patterns will change within all STG is uncertain at this point. However, an intensifying hydrological cycle is expected to increase precipitation in many areas (high confidence WGI 2.5, 14.2), although longer droughts are also expected in other STG (medium Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 confidence). Changes in the hydrological cycle impact on coastal ecosystems, increasing damage through coastal flooding and physical damage from storm waves [Mumby et al., 2011]. Improving our understanding of how weather systems associated with features such as the SPCZ (WGI 14.3.1) will vary is critical to climate change adaptation of a large number of nations associated with the STG. Developing an understanding of how ocean temperature, climate systems such as the SPCZ and ITCZ, and climate change and variability (e.g., ENSO, PDO) interact will be essential in this regard. For example, variability in the latitude of the SPCZ is projected to increase, possibly leading to more extreme events in Pacific island countries [Cai et al., 2012]. The consequences of projected sea temperatures on the frequency of coral bleaching and mortality within key sub- regions of the STG are outlined in Box CC-CR, Figures 30.10 and Figure SM30-3. As with other sub-regions (particularly CBS, STG, and SES) dominated by coral reefs, mass coral bleaching and mortality becomes an annual risk under all scenarios, with mass mortality events beginning to occur every 1 2 years by 2100 (virtually certain, Box CC-CR, Figure 30-10, Figure SM30-3). Coral-dominated reef ecosystems (areas with more than 30% coral cover) are very likely to disappear under these circumstances by the mid part of this century [van Hooidonk et al., 2013]. The loss of substantial coral communities has implications for the three-dimensional structure of coral reefs (Box CC-CR) and the role of the latter as habitat for organisms such as fish [Hoegh-Guldberg, 2011; Hoegh- Guldberg et al., 2011a; Pratchett et al., 2011a; Bell et al., 2013b]. The consequences of increasing sea temperature can be exacerbated by increasing ocean acidification, with potential implications for reef calcification ([Kleypas et al., 1999; Hoegh-Guldberg et al., 2007; Doney et al., 2009], medium confidence), reef metabolism and community calcification [Dove et al., 2013], and other key ecological processes ([Pörtner et al., 2001; Pörtner et al., 2007; Munday et al., 2009]. Ocean pH within the STG will continue to decrease as atmospheric CO2 increases, bringing pH within the STG to 7.9 and 7.7 at atmospheric concentrations of 450 ppm and 800 ppm, respectively (Figure SM30-2a, Box CC-OA). Aragonite saturation states will decrease to around 1.6 (800 ppm) and 3.3 (450 ppm; Figure SM30-2b). Decreasing carbonate ion concentrations and saturation states pose serious risks to other marine calcifiers such as encrusting coralline algae, coccolithophores (phytoplankton), and a range of benthic invertebrates [Doney et al., 2009; Feely et al., 2009]. Increasing sea temperatures and sea level are also likely to influence other coastal ecosystems (e.g., mangroves, seagrass meadows) in the Pacific, although significant gaps and uncertainties exist (29.3.4 [Waycott et al., 2007; Waycott et al., 2011]. Many of the negative consequences for coral reefs, mangroves, and seagrass meadows are likely to have negative consequences for dependent coastal fisheries (through destruction of habitat) and tourism industries (medium confidence) ([Bell et al., 2011b; Pratchett et al., 2011a; Pratchett et al., 2011b; Bell et al., 2013a]. Populations of key large pelagic fish are projected to move many hundreds of kilometers east of where they are today in the Pacific STG (high confidence) [Lehodey et al., 2008; Lehodey et al., 2010a; Lehodey et al., 2011; Lehodey et al., 2013], with implications for income, industry, and food security across multiple Pacific Island nations (high confidence)[Cheung et al., 2010; McIlgorm et al., 2010; Bell et al., 2011b; Bell et al., 2013a], 7.4.2, Table 29.2-29.3). These predictions of species range displacements, contractions, and expansions in response to anticipated changes in the Ocean (Box CC-MB) present both a challenge and an opportunity for the development of large-scale management strategies to preserve these valuable species. Our understanding of the consequences of reduced O2 for pelagic fish populations is not clear, although there is high agreement on the potential physiological outcomes (6.3.3). Those species that are intolerant to hypoxia, such as skipjack and yellowfin tuna [Lehodey et al., 2011], will have their depth range compressed in the Pacific STG, which will increase their vulnerability to fisheries and reduce overall fisheries habitat and productivity (medium confidence) [Stramma et al., 2010; Stramma et al., 2011]. Despite the importance of these potential changes, our understanding of the full range of consequences is limited at this point. 30.5.7. Deep Sea (>1000 m) Assessments of the influence of climate change on the Deep Sea (DS) are challenging due to difficulty of access and scarcity of long-term, comprehensive observations [Smith et al., 2009]. The size of this habitat is also vast, covering well over 54% of the earth s surface and stretching from the top of the mid-oceanic ridges to the bottom of deep ocean trenches [Smith et al., 2009]. The fossil record in marine sediments reveals that the DS has undergone large changes in response to climate change in the past [Knoll and Fischer, 2011]. The paleo-skeletal record shows it is Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 the rate, not just the magnitude, of climate change (temperature, O2, and CO2) that is critical to marine life in DS. The current rate of change in key parameters very likely exceeds that of other major events in Earth history. Two primary time scales are of interest. The first is the slow rate (century-scale) of ocean circulation and mixing, and consequently the slow rate at which DS ecosystems experience physical climate change. The second is the rapid rate at which organic matter enters the deep ocean from primary productivity generated at surface of the Ocean, which represents a critical food supply to DS animals [Smith and Kaufmann, 1999; Smith et al., 2009]. It can also represent a potential risk in some circumstances where the flux of organic carbon into the deep ocean, coupled with increased sea temperatures, can lead to anoxic areas (dead zones) as metabolism is increased and O2 decreased [Chan et al., 2008; Stramma et al., 2010]. 30.5.7.1. Observed Changes and Potential Impacts The greatest rate of change of temperature is occurring in the upper 700 m of the Ocean (WGI 3.2, very high confidence), although smaller yet significant changes are occurring at depth. The DS environment is typically cold (~-0.5 3°C; [Smith et al., 2008]), although abyssal temperatures in the SES can be higher (e.g., Mediterranean DS ~12°C, [Danovaro et al., 2010]. In the latter case, DS organisms can thrive in these environments as well, illustrating the variety of temperature conditions that differing species of abyssal life have adapted to. Individual species, however, are typically constrained within a narrow thermal and O2-demand window of tolerance [Pörtner, 2010] and therefore it is likely that shifts in the distribution of DS species and regional extinctions will occur. Warming over multiple decades has been observed below 700 m [Levitus et al., 2005; Levitus et al., 2009], with warming being minimal at mid-range depths (2000 3000 m), and increasing towards the sea floor in some sub- regions (e.g., Southern Ocean, WGI Chapter 3). For the deep Atlantic Ocean, the mean age of deep waters (mean time since last exposure to the atmosphere) is ~250 years; the oldest deep waters of the Pacific Ocean are >1000 years old. The patterns of ocean circulation are clearly revealed by the penetration of tracers and the signal of CO2 released from burning fossil-fuel penetrating into the abyss [Sabine et al., 2004]. It will take many centuries for full equilibration of deep ocean waters and their ecosystems with recent planetary warming and CO2 levels [Wunsch and Heimbach, 2008]. Temperature accounts for ~86% of the variance in the export of organic matter to the DS (medium confidence) [Laws et al., 2000]. Consequently, upper ocean warming will reduce the export of organic matter to the DS (medium confidence), potentially changing the distribution and abundance of DS organisms and associated food webs, and ecosystem processes [Smith and Kaufmann, 1999]. Most organic matter entering the DS is recycled by microbial systems at relatively shallow depths [Buesseler et al., 2007], at rates that are temperature dependent. Upper ocean warming will increase the rate of sub-surface decomposition of organic matter (high confidence), thus intensifying the intermediate depth oxygen minimum zones [Stramma et al., 2008; Stramma et al., 2010] and reducing food supply to the abyssal ocean. Particulate organic carbon is exported from the surface to deeper layers of the Ocean (>500m) with an efficiency of between 20 50% [Buesseler et al., 2007], much of it being recycled by microbes before it reaches 1000m [Smith et al., 2009]. The export of organic carbon is dependent on surface net primary productivity, which is likely to vary (Box CC-PP), influencing the supply of food to DS [Laws et al., 2000; Smith et al., 2008]. Warming of intermediate waters will also increase respiration at mid-water depths, reducing the flux of organic carbon. Our understanding of other components of DS ecosystems is also relatively poor. For example, there is limited evidence and limited agreement as to how ocean warming and acidification are likely to affect ecosystems such as those associated with hydrothermal vents [Van Dover, 2012]. Oxygen concentrations are decreasing in the DS [Stramma et al., 2008; Helm et al., 2011a]. Although, the largest signals occur at intermediate water depths < 1000 m [Nakanowatari et al., 2007; Whitney et al., 2007; Falkowski et al., 2011], some waters >1000 m depth are also experiencing a decline [Jenkins, 2008]. The quantity of dissolved O2 throughout the Ocean will be reduced with warming due to direct effects on solubility (high confidence), with these effects being widely distributed [Shaffer et al., 2009]. It is also virtually certain that metabolic rates of all animals and microbial respiration rates will increase with temperature [Brown et al., 2004]. Thus, increased microbial activity and reduced O2 solubility at higher temperatures will have additive consequences for the decline of O2 (high Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 confidence) even in the DS. The DS waters are relatively well-oxygenated due to the higher solubility of O2 in colder waters and the low supply rate of organic matter to great depths. The availability of oxygen to marine animals is governed by a combination of concentration, temperature, pressure, and related properties such as diffusivity. Analysis by [Hofmann et al., 2013] reveals that the supply potential of oxygen to marine animals in cold deep waters is similar to that at much shallower depths (very high confidence). Anthropogenic CO2 has penetrated to at least 1000 m in all three ocean basins (particularly the Atlantic; [Doney et al., 2009]. Further declines of calcite and aragonite in already under-saturated DS water will presumably decrease biological carbonate structure formation and increase dissolution, as has happened many times in Earth s past (high confidence) [Zeebe and Ridgwell, 2011]. Some cold-water corals (reported down to 3500m) already exist in waters under-saturated with respect to aragonite [Lundsten et al., 2009]. While initial investigations suggested that ocean acidification (reduced by 0.15 and 0.30 pH units) would result in a reduction in the calcification rate of deep water corals (30% and 6%, respectively [Maier et al., 2009]), there is accumulating evidence that ocean acidification may have far less impact than previously anticipated on the calcification of some deep water corals (limited evidence, medium agreement, low confidence) although it may reduce important habitats given that dead unprotected coral mounds are likely to dissolve in under-saturated waters [Thresher et al., 2011; Form and Riebesell, 2012; Maier et al., 2013]. 30.5.7.2. Key Risks and Vulnerabilities Rising atmospheric CO2 poses a risk to DS communities through increasing temperature, and decreasing O2, carbonate chemistry and pH (high confidence) [Keeling et al., 2010]. Risks associated with the DS have implications for the Ocean and planet given the high degree of inherent dependency and connectivity. The resulting changes to the flow of organic carbon to some parts of the DS (e.g., STG) are very likely to affect DS ecosystems (medium confidence) [Smith et al., 2008]. As with the Ocean generally, there is a need to fill in the substantial gaps that exist in our knowledge and understanding of the world s largest habitat and its responses to rapid anthropogenic climate change. 30.5.8. Detection and Attribution of Climate Change Impacts with Confidence Levels The analysis in Chapter 30 and elsewhere in AR5 has identified a wide range of physical, chemical, and ecological components that have changed over the last century (Box CC-MB). Figure 30-11 summarizes a number of examples from the Ocean as a region together with the degree of confidence in both the detection and attribution steps. For ocean warming and acidification, confidence is very high that changes are being detected and that they are due to changes to the atmospheric greenhouse gas content. There is considerable confidence in both the detection (very high confidence) and attribution (high confidence) of mass coral bleaching and mortality (Figure 30-11b), given the well-developed understanding of environmental processes and physiological responses driving these events (Box CC-CR, 6.3.1). For other changes, confidence is lower, either because detection of changes has been difficult, or monitoring programs are not long-established (e.g., field evidence of declining calcification) or because detection has been possible but models are in conflict (e.g., wind-driven upwelling). The detection and attribution of recent changes is discussed in further detail in (18.3.3 4). [INSERT FIGURE 30-11 HERE Figure 30-11: Expert assessment of degree of confidence in detection and attribution of physical and chemical changes (a) and ecological changes (b) across sub-regions, as designated in Figure 30-1a, and processes in the Ocean (based on evidence explored throughout Chapter 30 and elsewhere in AR5). Further explanation of this figure is given in 18.3.3 4 and 18.6.] Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.6. Sectorial Impacts, Adaptation, and Mitigation Responses Human welfare is highly dependent on ecosystem services provided by the Ocean. Many of these services are provided by coastal and shelf areas, and are consequently addressed in other chapters (e.g., 5.4.3, 7.3.2.4, 22.3.2.3) Oceans contribute provisioning (e.g., food, raw materials; see 30.6.4.1), regulating (e.g., gas exchange, nutrient recycling, carbon storage, climate regulation, water flux), supporting (e.g., habitat, genetic diversity) and cultural (e.g., recreational, cultural) services [Millennium-Ecosystem-Assessment, 2005; Tallis et al., 2013]. The accumulating evidence indicating that fundamental ecosystem services within the Ocean are shifting rapidly should be of major concern, especially with respect to the ability of regulating and supporting ecosystem services to underpin current and future human population demands [Rockström et al., 2009; Ruckelshaus et al., 2013]. Discussion here is restricted to environmental, economic, and social sectors that have direct relevance to the Ocean, namely natural ecosystems, fisheries and aquaculture, tourism, shipping, oil and gas, human health, maritime security, and renewable energy. The influences of climate change on Ocean sectors will be mediated through simultaneous changes in multiple environmental and ecological variables (Figure 30-12), and the extent to which changes can be adapted to and/or risks mitigated (Table 30-3). Both short-term and longer-term adaptation is necessary to address impacts arising from warming, even under the lowest stabilization scenarios assessed. [INSERT FIGURE 30-12 HERE Figure 30-12: (a) Examples of projected impacts and vulnerabilities associated with climate change in Ocean sub- regions. (b) Examples of risks to fisheries from observed and projected impacts across Ocean sub-regions. Letters indicate level of confidence: (vL): Very low, (L): Low, (M): Medium, (H): High and (vH): Very high. Details of sub-regions are given in Table 30-1a and 30.1.1. Sectorial approaches dominate resource management in the Ocean (e.g., shipping tends to be treated in isolation from fishing within an area), yet cumulative and interactive effects of individual stressors are known to be ubiquitous and substantial [Crain et al., 2008]. Climate change consistently emerges as a dominant stressor in regional to global-scale assessments, although land-based pollution, commercial fishing, invasive species, coastal habitat modification, and commercial activities such as shipping all rank high in many places around the world (e.g., 30.5.3, 30.5.4, 5.3.4) [Halpern et al., 2009; Halpern et al., 2010]. Such cumulative effects pose challenges to managing for the full suite of stressors to marine systems, but also present opportunities where mitigating a few key stressors can potentially improve overall ecosystem condition (e.g., [Halpern et al., 2010; Kelly et al., 2011]. The latter has often been seen as a potential strategy for reducing negative consequences of climate impacts on marine ecosystems by boosting ecosystem resilience, thus buying time while the core issue of reducing greenhouse gas emissions is tackled [West et al., 2009]. 30.6.1. Natural Ecosystems Adaptation in natural ecosystems may occur autonomously, such as shifts in species composition and distributions [Poloczanska et al., 2013] or engineered by human intervention, such as assisted dispersal (4.4, [Hoegh-Guldberg et al., 2008]. Currently, adaptation strategies for marine ecosystems included reducing additional stressors (e.g. maintaining water quality, adapting fisheries management) and maintaining resilience ecosystems (e.g., Marine Protected Areas) and are moving towards whole-of-ecosystem management approaches. Coral reefs, for example, will recover faster from mass coral bleaching and mortality if healthy populations of herbivorous fish are maintained (medium confidence, [Hughes et al., 2003], indicating that reducing overfishing will help maintain coral-dominated reef systems while the international community reduces the emissions of greenhouse gases to stabilize global temperature and ocean chemistry. Approaches such as providing a formal valuation of ecological services from the Ocean have the potential to facilitate adaptation by underpinning more effective governance, regulation, and ocean policy while at the same time potentially improving the management of these often vulnerable services through the development of market mechanisms and incentives [Beaudoin and Pendleton, 2012]. Supporting, regulating, and cultural ecosystem services tend to transcend the immediate demands placed on provisioning services and are difficult to value in Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 formal economic terms due to their complexity, problems such as double counting, and the value of non-market goods and services arising from marine ecosystems generally [Fu et al., 2011; Beaudoin and Pendleton, 2012]. Blue Carbon is defined as the organic carbon sequestered by marine ecosystems such as phytoplankton, mangrove, seagrass, and salt marsh ecosystems [Laffoley and Grimsditch, 2009; Nellemann and Corcoran, 2009]. In this respect, Blue Carbon will provide opportunities for both adaptation to, and mitigation of, climate change if key uncertainties in inventories, methodologies, and policies for measuring, valuing, and implementing Blue Carbon strategies are resolved [McLeod et al., 2011]. Sediment surface levels in vegetated coastal habitats can rise several meters over thousands of years, building carbon-rich deposits [Brevik and Homburg, 2004; Lo Iacono et al., 2008]. The degradation of coastal habitats not only liberates much of the carbon associated with vegetation loss, but can release and oxidize buried organic carbon through erosion of cleared coastlines (high confidence) [Duarte et al., 2005]. Combining data on global area, land-use conversion rates, and near-surface carbon stocks for marshes, mangroves, and seagrass meadows, Pendleton et al. [2012] revealed that the CO2 emissions arising from destruction of these three ecosystems was equivalent to 3 19% of the emissions generated by deforestation globally, with economic damages estimated to be US$6 42 billion annually. Similarly, Luisetti et al. [2013] estimate the carbon stock of seagrass and salt marshes in Europe, representing <4% of global carbon stocks in coastal vegetation, was valued at US$180 million, at EU Allowance price of 8/tCO2 in June 2012. A reversal of EU Environmental Protection Directives could result in economic losses of US$1 billion by 2060. Blue Carbon strategies can also be justified in light of the numerous ecosystem services these ecosystems provide, such as protection against coastal erosion and storm damage, and provision of habitats for fisheries species (17.4). 30.6.2. Economic Sectors 30.6.2.1. Fisheries and Aquaculture The Ocean provided 64% of the production (in tonnes) supplied by world fisheries (capture and aquaculture) in 2010, amounting to 148.5 million tonnes of fish and shellfish [FAO, 2012]. This production, valued at US$217.5 billion, and supplied, on average, 18.6 kg of protein-rich food per person to an estimated population of 6.9 billion [FAO, 2012]. Marine capture fisheries supplied 77.4 million tonnes with highest production from the northwest Pacific (27%), west-central Pacific (15%), northeast Atlantic (11%) and southeast Pacific (10%)[FAO, 2012]. World aquaculture production (59.9 million tonnes in 2010) is dominated by freshwater fishes, nevertheless marine aquaculture supplied 18.1 million tonnes (30%)[FAO, 2012]. Marine capture fisheries production increased from 16.8 million tonnes in 1950 to a peak of 86.4 million tonnes in 1996, then declined before stabilising around 80 million tonnes [FAO, 2012]. The stagnation of marine capture fisheries production is attributed to full exploitation of around 60% of the world s marine fisheries and overexploitation of 30% (estimates for 2009) [FAO, 2012]. Major issues for industrial fisheries include illegal, unreported and unregulated fishing, ineffective implementation of monitoring, control and surveillance, and overcapacity in fishing fleets [Bank and FAO, 2008; FAO, 2012]. Such problems are being progressively addressed in several developed and developing countries [Hilborn, 2007; Pitcher et al., 2009; Worm et al., 2009], where investments have been made in stock assessment, strong management, and application of the FAO Code of Conduct for Responsible Fisheries and the FAO Ecosystem Approach to Fisheries Management. The significance of marine capture fisheries is illustrated powerfully by the number of people engaged in marine small-scale fisheries (SSF) in developing countries. SSF account for around half of the fish harvested from the Ocean, and provide jobs for more than 47 million people about 12.5 million fishers and another 34.5 million people engaged in post-harvest activities [Mills et al., 2011]. SSF are often characterized by large numbers of politically-weak fishers operating from decentralized localities, with poor governance, and insufficient data to monitor catches effectively [Kurien and Willmann, 2009; Cochrane et al., 2011; Pomeroy and Andrew, 2011]. For these SSF, management that aims to avoid further depletion of overfished stocks may be more appropriate in the short-term than management aimed at maximizing sustainable production. These aims are achieved through adaptive management by: (1) introduction of harvest controls (e.g., size limits, closed seasons and areas, gear restrictions, and protection of spawning aggregations) to avoid irreversible damage to stocks in the face of uncertainty [Cochrane et Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 al., 2011]; (2) flexible modification of these controls through monitoring [Plagányi et al., 2013]; and (3) investing in the social capital and institutions needed for communities and governments to manage SSF [Makino et al., 2009; Pomeroy and Andrew, 2011]. Changes to ocean temperature, chemistry, and other factors are generating new challenges for fisheries resulting in loss of coastal and oceanic habitat [Hazen et al., 2012; Stramma et al., 2012], the movement of species [Cheung et al., 2011], the spread and increase of disease and invading species [Ling, 2008; Raitsos et al., 2010; Chan et al., 2011], and changes in primary production[Chassot et al., 2010]. There is medium evidence and medium agreement that these changes will change both the nature of fisheries and their ability to provide food and protein for hundreds of millions of people (7.2.1.2). The risks to ecosystems and fisheries vary from region to region (7.3.2.4). Dynamic bioclimatic envelope models under SRES A1B project potential increases in fisheries production high latitudes, and potential decreases at lower latitudes by the mid-21st century [Cheung et al., 2010] (6.5). Overall, warming temperatures are projected to shift optimal environments for individual species polewards and redistribute production, however changes will be region specific [Cheung et al., 2010; Merino et al., 2012]. Fisheries, in particular shellfish, are also vulnerable to declining pH and carbonate ion concentrations. As a result, the global production of shellfish fisheries is likely to decrease, [Cooley and Doney, 2009; Pickering et al., 2011]) with further ocean acidification (medium confidence) (6.3.2, 6.3.5, 6.4.1.1, Box CC-OA). Impacts may be first observed in EBUE where upwelled water is already relatively low in O2 and undersaturated with aragonite (30.5.5). Seasonal upwelling of acidified waters onto the continental shelf in the California Current region, has recently affected oyster hatcheries along the coast of Washington and Oregon (30.5.5.1.1[Barton et al., 2012]). Whether declining pH and aragonite saturation due to climate change played a role is unclear, however future declines will increase the risk of such events occurring. Most marine aquaculture species are sensitive to changing ocean temperature (6.3.1.4, exposed through pens, cages and racks placed directly in the sea, utilization of seawater in land-based tanks or collection of wild spat) and, for mollusks particularly, changes in carbonate chemistry (6.3.2.4 [Turley et al., 2011; Barton et al., 2012]). Environmental changes can therefore impact farm profitability, depending on target species and farm location. For example, a 1°C rise in SST is projected to shift production of Norwegian salmonids further north but may increase production overall [Hermansen and Heen, 2012]. Industries for non-food products, which can be important for regional livelihoods such as Black Pearl in Polynesia, are also affected by rising SST. Higher temperatures are known to affect the quality of pearl nacre, and can increase levels of disease in adult oysters [Bell et al., 2011a; Pickering et al., 2011; Bell et al., 2013b]. Aquaculture production is also vulnerable to extreme events such as storms and floods (e.g., [Chang et al., 2013]). Flooding and inundation by seawater may be a problem to shore facilities on low-lying coasts. For example, shrimp farming operations in the tropics will be challenged by rising sea levels, which will be exacerbated by mangrove encroachment and reduce the ability for thorough-drying of ponds between crops [Della Patrona et al., 2011]. The impacts of climate change on marine fish stocks are expected to affect the economics of fishing and livelihoods in fishing nations through changes in the price and value of catches, fishing costs, income to fishers and fishing companies, national labor markets, and industry re-organization [Sumaila et al., 2011] (6.4.1). A study of the potential vulnerabilities of national economies to the effects of climate change on fisheries, in terms of exposure to warming, relative importance of fisheries to national economies and diets, and limited societal capacity to adapt, concluded that a number countries including Malawi, Guinea, Senegal, Uganda, Sierra Leone, Mozambique, Tanzania, Peru, Columbia, Venezuela, Mauritania, Morocco, Bangladesh, Cambodia, Pakistan, Yemen, and Ukraine are most vulnerable [Allison et al., 2009]. Aquaculture production is expanding rapidly [Bostock et al., 2010] and will play an important role in food production and livelihoods as the human demand for protein grows. This may also add pressure on capture fisheries (7.3.2.6) [FAO, 2012; Merino et al., 2012]. Two-thirds of farmed food fish production (marine and freshwater) is achieved with the use of feed derived from wild-harvested, small, pelagic fish and shellfish. Fluctuations in the availability and price of fishmeal and fish oil for feeds, as well as their availability, pose challenges for the growth of sustainable aquaculture production, particularly given uncertainties in changes in EBUE upwelling dynamics to climate change (30.5.5). Technological advances and management change such as increasing feed efficiencies, using Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 alternatives to fishmeal and fish oil, and farming of herbivorous finfish, coupled with economic and regulatory incentives will reduce the vulnerability of aquaculture to the impacts of climate change on small, pelagic fish abundance [Naylor et al., 2009; Merino et al., 2010; FAO, 2012]. The challenges of optimizing the economic and social benefits of both industrial fisheries, SSF and aquaculture operations, which often already include strategies to adapt to climatic variability [Salinger et al., 2013], are now made more complex by climate change [Cochrane et al., 2009; Brander, 2010; 2013]. Nevertheless, adaptation options include establishment of early-warning systems to aid decision-making, diversification of enterprises and development of adaptable management systems [Chang et al., 2013]. Vulnerability assessments that link oceanographic, biological, and socio-economic systems can be applied to identify practical adaptations to assist enterprises, communities, and households to reduce the risks from climate change and capitalize on the opportunists [Pecl et al., 2009; Bell et al., 2013b; Norman-López et al., 2013]. The diversity of these adaptation options, and the policies needed to support them, are illustrated by the following examples. 30.6.2.1.1. Tropical fisheries based on large pelagic fish Fisheries for skipjack, yellowfin, big-eye, and albacore tuna provide substantial economic and social benefits to the people of Small Island Developing States (SIDS). For example, tuna fishing license fees contribute substantially (up to 40%) to the government revenue of several Pacific Island nations [Gillett, 2009; Bell et al., 2013b]. Tuna fishing and processing operations also contribute up to 25% of gross domestic product in some of these nations and employ over 12,000 people [Gillett, 2009; Bell et al., 2013b]. Considerable economic benefits are also derived from fisheries for top pelagic predators in the Indian and Atlantic Oceans [FAO, 2012; Bell et al., 2013a]. Increasing sea temperatures and changing patterns of upwelling are projected to cause shifts in the distribution and abundance of pelagic top predator fish stocks (30.5.2, 30.5.5, 30.5.6), with potential to create winners and losers among island economies as catches of the trans-boundary tuna stocks change among and within their exclusive economic zones (EEZs;[Bell et al., 2013b; Bell et al., 2013a]. A number of practical adaptation options and supporting policies have been identified to minimize the risks and maximize the opportunities associated with the projected changes in distribution of the abundant skipjack tuna in the tropical Pacific ([Bell et al., 2011; Bell et al., 2013a], Table 30-2). These adaptation and policy options include: (1) full implementation of the regional vessel day scheme , designed to distribute the economic benefits from the resource in the face of climatic variability, and other schemes to control fishing effort in subtropical areas; (2) strategies for diversifying the supply of fish for canneries in the west of the region as tuna move progressively east; (3) continued effective fisheries management of all tuna species; (4) energy efficiency programs to assist domestic fleets to cope with increasing fuel costs and the possible need to fish further from port; and (5) the eventual restructuring of regional fisheries management organizations to help coordinate management measures across the entire tropical Pacific. Provision of operational-level catch and effort data from all industrial fishing operations to improve models for projecting redistribution of tuna stocks and quotas under climate change [Salinger et al. 2013][Nicol et al., 2013]. Similar adaptation options and policy responses are expected to be relevant to the challenges faced by tuna fisheries in the tropical and sub-tropical Indian and Atlantic Oceans. [INSERT TABLE 30-2 HERE Table 30-2: Examples of priority adaptation options and supporting policies to assist Pacific Island countries and territories to minimize the threats of climate change to the socio-economic benefits derived from pelagic and coastal fisheries and aquaculture, and to maximize the opportunities. These measures are classified as win-win (W-W) adaptations, which address other drivers of the sector in the short-term and climate change in the long-term, or lose- win (L-W) adaptations, where benefits exceed costs in the short-term but accrue under longer-term climate change (modified from [Bell et al., 2013b]). Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.6.2.1.2. Small-scale fisheries Small-scale fisheries (SSF) account for 56% of catch and 91% of people working in fisheries in developing countries [Mills et al., 2011]. SSF are fisheries that tend to operate at family or community level, have low levels of capitalization, and make an important contribution to food security and livelihoods. They are often dependent on coastal ecosystems, such as coral reefs, that provide habitats for a wide range of harvested fish and invertebrate species. Despite their importance to many developing countries, such ecosystems are under serious pressure from human activities including deteriorating coastal water quality, sedimentation, ocean warming, overfishing, and acidification (7.2.1.2, 30.3, 30.5, Box CC-CR). These pressures are translating into a steady decline in live coral cover, which is very likely to continue over the coming decades, even where integrated coastal zone management is in place (30.5.4, 30.5.6. For example, coral losses around Pacific Islands are projected to be as high as 75% by 2050 [Hoegh-Guldberg et al., 2011a]. Even under the most optimistic projections (a 50% loss of coral by 2050), changes to state of coral reefs (Box CC-CR, Figure 30-10, Figure 30-12) are very likely to reduce the availability of associated fish and invertebrates that support many of the SSF in the tropics (high confidence). In the Pacific, the productivity of SSF on coral reefs has been projected to decrease by at least 20% by 2050 [Pratchett et al., 2011b], which is also likely to occur in other coral reef areas globally given the similar and growing stresses in these other regions (Table SM30-1, 30.5.4). Adaptation options and policies for building the resilience of coral reef fisheries to climate change suggested for the tropical Pacific include: (1) strengthening the management of catchment vegetation to improve water quality along coastlines; (2) reducing direct damage to coral reefs; (3) maintaining connectivity of coral reefs with mangrove and seagrass habitats; (4) sustaining and diversifying the catch of coral reef fish to maintain their replenishment potential; and (5) transferring fishing effort from coral reefs to skipjack and yellowfin tuna resources by installing anchored fish-aggregating devices (FADs) close to shore [Bell et al., 2011b; Bell et al., 2013b; Bell et al., 2013a], Table 30-2). These adaptation options and policies represent a no regrets strategy in that they provide benefits for coral reef fisheries and fishers irrespective of climate change and ocean acidification. 30.6.2.1.3. Northern Hemisphere HLSBS fisheries The high latitude fisheries in the northern hemisphere span from around 30/35°N to 60°N in the North Pacific and 80°N in the North Atlantic, covering a wide range of thermal habitats supporting subtropical/temperate species to boreal/arctic species. The characteristics of these HLSBS environments, as well as warming trends, are outlined in 30.5.1 and Table 30-1. In part, as a result of 30 years of increase in temperature [Belkin, 2009; Sherman et al., 2009], there has been an increase in the size of fish stocks associated with high latitude fisheries in the northern hemisphere. This is particularly the case for the Norwegian spring-spawning herring, which has recovered from near-extinction as a result of overfishing and a cooler climate during the 1960s [Toresen and Ostvedt, 2000]. The major components of both pelagic and demersal high latitude fish stocks are boreal species located north of 50°N. Climate change is projected to increase high latitude plankton production and displace zooplankton and fish species poleward. As a combined result of these future changes, the abundance of fish (particularly boreal species) may increase in the northernmost part of the high latitude region [Cheung et al., 2011], although increases will only be moderate in some areas. The changes in distribution and migration of the pelagic fishes shows considerable spatial and temporal variability, which can increase tensions among fishing nations. In this regard, tension over the Atlantic mackerel fisheries has led to what many consider the first climate-change related conflict between fishing nations ([Cheung et al., 2012], 30.6.5), and which has emphasized the importance of developing international collaboration and frameworks for decision making [Miller et al., 2013](30.6.7, 15.4.3.3). The Atlantic mackerel has over the recent decades been a shared stock between the EU and Norway. However, the recent advancement of the Atlantic mackerel into the Icelandic EEZ during summer has resulted in Icelandic fishers operating outside the agreement between the EU and Norway. Earlier records of mackerel from the first half of the 20th and second half of the 19th century show, however, that mackerel was present in Icelandic waters during the earlier warm periods [Astthorsson et al., 2012]. In the Barents Sea, the North-east Arctic cod, Gadus morhua, reached record-high abundance during 2012 and also reached its northernmost-recorded distribution (82°N)[ICES, 2012]. A further northward migration is impossible Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 since this would be into the Deep Sea Polar Basin, beyond the habitat of shelf species. A further advancement eastwards to the Siberian shelf is, however, possible. The North-east Arctic cod stock is shared exclusively by Norway and Russia, and to date there has been a good agreement between those two nations on the management of the stock. These examples highlight the importance of international agreements and cooperation (Table 30-4). The HLSBS fisheries constitute a large-scale high-tech industry, with large investments in highly mobile fishing vessels, equipment, and land-based industries with capacity for adapting fisheries management and industries for climate change [Frontiers-Economics-Ltd, 2013]. Knowledge of how climate fluctuations and change affect the growth, recruitment, and distribution of fish stocks is presently not incorporated into fisheries management strategies [Perry et al., 2010]. These strategies are vital for fisheries that hope to cope with the challenges of a changing ocean environment, and are centrally important to any attempt to develop ecosystem-based management and sustainable fisheries under climate change. The large pelagic stocks, with their climate-dependent migration pattern, are shared among several nations. Developing equitable sharing of fish quotas through international treaties (Table 30-4) is a necessary adaptation for a sustainable fishery. Factors presently taken into account in determining the shares of quotas are the historical fishery, bilateral exchanges of quotas for various species, and occupation time of the stocks in the various EEZs. 30.6.2.2. Tourism Tourism recreation represents one of the world s largest industries, accounting for 9% (>US$6 trillion) of global GDP and employing over 255 million people. It is expected to grow by an average of 4% annually and reach 10% of global GDP within the next 10 years [WTTC, 2012]. As with all tourism, that which is associated with the Ocean is heavily influenced by climate change, global economic and socio-political conditions, and their interactions ([Scott et al., 2012b]; 10.6.1). Climate change, through impacts on ecosystems (e.g., coral reef bleaching), can reduce the appeal of destinations, increase operating costs, and/or increase uncertainty in a highly sensitive business environment [Scott et al., 2012b]. Several facets of the influence of climate change on the Ocean directly impact tourism (10.6.1, 10.6.2). Tourism is susceptible to extreme events such as violent storms, long periods of drought, and/or extreme precipitation events (5.3.3, 10.6.1, [IPCC, 2012]). Sea level rise through its influence on coastal erosion and submergence, salinization of water supplies, and changes to storm surge, increases the vulnerability of coastal tourism infrastructure, tourist safety, and iconic ecosystems (high confidence) (5.3.3.2 10.6.1, [IPCC, 2012], Table SPM.1). For example, approximately 29% of resorts in the Caribbean are within 1 m of the high tide mark and 60% are at risk of beach erosion from rapid sea level rise [Scott et al., 2012a]. Increasing sea temperatures (30.3.1.1) can change the attractiveness of locations and the opportunities for tourism through their influence on the movement of organisms and the state of ecosystems such as coral reefs (10.6.2, Box CC-CR, [UNWTO and UNEP, 2008]). Mass coral bleaching and mortality (triggered by elevated sea temperatures, high confidence) can decrease the appeal of destinations for diving-related tourism, although the level of awareness of tourists of impacts (e.g., <50% of tourists were concerned about coral bleaching during 1998) and expected economic impacts have been found to be uncertain [Scott et al., 2012b]. Some studies, however, have noted reduced tourist satisfaction and identified dead coral as one of the reasons for disappointment at the end of the holiday [Westmacott et al., 2001]. Tourists respond to changes in factors such as weather and opportunity by expressing different preferences. For example, preferred conditions and hence tourism are projected to shift towards higher latitudes with climate change, or from summer to cooler seasons [Amelung et al., 2007] (10.6.2). Options for adaptation by the marine tourism sector include: (1) identifying and responding to inundation risks with current infrastructure, and planning for projected sea level rise when building new tourism infrastructure (5.5, [Scott et al., 2012a]); (2) promoting shoreline stability and natural barriers by preserving ecosystems such as mangroves, salt marshes and coral reefs (5.5, [Scott et al., 2012b]); (3) deploying forecasting and early-warning systems in order to anticipate challenges to tourism and natural ecosystems [Strong et al., 2011; IPCC, 2012]; (4) preparation of risk management and disaster preparation plans in order to respond to extreme events; (5) reducing the effect of other stressors on ecosystems and building resilience in iconic tourism features such as coral reefs and mangroves; and (6) Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 educating tourists to improve understanding the negative consequences of climate change over those stemming from local stresses [Scott et al., 2012b; Scott et al., 2012a]. Adaptation plans for tourism industries need to address specific operators and regions. For example, some operators may have costly infrastructure at risk while others may have few assets but are dependent on the integrity of natural environments [Turton et al., 2010]. 30.6.2.3. Shipping International shipping accounts for >80% of world trade by volume [UNCTAD, 2009a; b] and ~3% of global CO2 emissions from fuel combustion although CO2 emissions are expected to increase 2-3 fold by 2050 [Heitmann and Khalilian, 2010], WGIII 8.1, 8.2). Changes in shipping routes [Borgerson, 2008], variation in the transport network due to shifts in grain production and global markets, as well as new fuel and weather-monitoring technology, may alter these emission patterns (WGIII 8.3, 8.5). Extreme weather events, intensified by climate change, may interrupt ports and transport routes more frequently, damaging infrastructure and introducing additional dangers to ships, crews, and the environment [UNCTAD, 2009a; b; Pinnegar et al., 2012](10.4.4). These issues have been assessed by some countries which have raised concerns over the potential for costly delays and cancellation of services, and the implications for insurance premiums as storminess and other factors change increase risks [Thornes et al., 2012]. Climate change may benefit maritime transport by reducing Arctic sea ice and consequently shorten travel distances between key ports [Borgerson, 2008] thus also decreasing total GHG emissions from ships (WGIII 8.5.2). Currently, reliability of this route limits its use [Schyen and Brathen, 2011], and the potential full operation of the Northwest Passage and Northern Sea Route would require a transit management regime, regulation (e.g., navigation, environmental, safety, and security), and a clear legal framework to address potential territorial claims that may arise, with a number of countries having direct interest in the Arctic. Further discussion of issues around melting Arctic sea ice and the Northern Sea Route are given in Chapter 28 (28.2.5, 28.3.4). 30.6.2.4. Offshore Energy and Mineral Resource Extraction and Supply The marine oil and gas industry face potential impacts from climate change on its ocean-based activities. Over 100 oil and gas platforms were destroyed in the Gulf of Mexico by the unusually strong hurricanes Katrina and Rita in 2005. Other consequences for oil pipelines and production facilities ultimately reduced US refining capacity by 20% [IPCC, 2012]. The increasing demand for oil and gas has pushed operations to waters 2000 m deep or more, far beyond continental shelves. The very large-scale moored developments required are exposed to greater hazards and higher risks, most of which are not well understood by existing climate/weather projections. Although there is a strong trend towards seafloor well completions with a complex of wells, manifolds, and pipes that are not exposed to surface forcing, these systems face different hazards from instability and scouring of the unconsolidated sediments by DS currents [Randolph et al., 2010]. The influence of warming oceans on sea floor stability is widely debated due largely to uncertainties about the effects of methane and methane hydrates [Sultan et al., 2004; Archer et al., 2009; Geresi et al., 2009]. Declining sea ice is also opening up the Arctic to further oil and gas extraction., Discussion of the potential expansion of oil and mineral production in the Arctic is made in Chapter 28 (28.2.5, 28.2.6, 28.3.4). The principal threat to oil and gas extraction and infrastructure in maritime settings is the impact of extreme weather [Kessler et al., 2011], which is likely to increase given that future storm systems are expected to have greater energy [Emanuel, 2005; Trenberth and Shea, 2006; Knutson et al., 2010]. Events such as Hurricane Katrina have illustrated challenges which will arise for this industry with projected increases in storm intensity [Cruz and Krausmann, 2008]. In this regard, early warning systems and integrated planning offer some potential to reduce the effect of extreme events [IPCC, 2012]. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.6.3. Human Health The major threats to public health due to climate change include diminished security of water and food supplies, extreme weather events, and changes in the distribution and severity of diseases, including those due to marine biotoxins ([Costello et al., 2009], 5.4.3.5, 6.4.2.3, 11.2). The predominately negative impacts of disease for human communities are expected to be more serious in low-income areas such as South-east Asia, southern and east Africa, and various sub-regions of South America [Patz et al., 2005], which also have under-resourced health systems [Costello et al., 2009]. Many of the influences are directly or indirectly related to basin-scale changes in the Ocean (e.g., temperature, rainfall, plankton populations, sea level rise, and ocean circulation [McMichael et al., 2006]). Climate change in the Ocean may influence the distribution of diseases like cholera (11.5.2.1), and the distribution and occurrence of harmful algal blooms (HAB). The frequency of cholera outbreaks induced by Vibrio cholerae and other enteric pathogens are correlated with sea surface temperatures, multidecadal fluctuations of ENSO, and plankton blooms, which may provide insight into how this disease may change with projected rates of ocean warming [Colwell, 1996; Pascual et al., 2000; Rodó et al., 2002; Patz et al., 2005; Myers and Patz, 2009; Baker- Austin et al., 2012]. The incidence of diseases such as ciguatera also shows a link to ENSO, with ciguatera becoming more prominent after periods of elevated sea temperature. This indicates that ciguatera may become more frequent in a warmer climate [Llewellyn, 2010], particularly given the higher prevalence of ciguatera in areas with degraded coral reefs (low confidence) [Pratchett et al., 2011a]. 30.6.4. Ocean-based Mitigation 30.6.4.1 Deep Sea Carbon Sequestration Carbon dioxide capture and storage into the deep sea and geologic structures are also discussed in WGIII Chapter 7 (7.5.5, 7.8.2, 7.10, 7.12) The economic impact of deliberate CO2 sequestration beneath the sea floor has previously been reviewed [IPCC, 2005]. Active CO2 sequestration from co-produced CO2 into sub-sea geologic formations is being instigated in the North Sea and in the Santos Basin offshore from Brazil. These activities will increase as offshore oil and gas production increasingly exploits fields with high CO2 in the source gas and oil. Significant risks from the injection of high levels of CO2 into deep ocean waters have been identified for DS organisms and ecosystems although chronic effects have not yet been studied. These risks are similar to those discussed previously with respect to ocean acidification and could further exacerbate declining O2 levels and changing trophic networks in deep water areas [Seibel and Walsh, 2001] (6.4.2.2). There are significant issues within the decision frameworks regulating these activities. Dumping of any waste or other matter in the sea, including the seabed and its subsoil, is strictly prohibited under the 1996 London Protocol (LP) except for those few materials listed in Annex I. Annex 1 was amended in 2006 to permit storage of CO2 under the seabed. Specific Guidelines for Assessment of Carbon Dioxide Streams for Disposal into Sub-Seabed Geological Formations were adopted by the parties to the LP in 2007. The Guidelines take a precautionary approach to the process, requiring Contracting Parties under whose jurisdiction or control such activities are conducted to issue a permit for the disposal subject to stringent conditions being fulfilled [Rayfuse and Warner, 2012]. 30.6.4.2 Offshore renewable energy Renewable energy supply from the Ocean includes ocean energy and offshore wind turbines. The global technical potential for ocean and wind energy is not as high as solar energy although considerable potential still remains. Detailed discussion of the potential of renewable energy sources are given in WGIII Chapter 7 (7.4.2, 7.5.3, 7.8.2). There is an increasing trend in the renewable energy sector to offshore wind turbines (10.2.2). At present, there is high uncertainty about how changes in wind intensity and patterns, and extreme events, will impact the offshore wind energy sector. Given the design and engineering solutions available to combat climate change impacts (Table 10.1, Table 10.7), it is unlikely that this sector will face insurmountable challenges from climate change. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.6.5. Maritime Security and Related Operations Climate change and its influence on the Ocean has become an area of increasing concern in terms of the maintenance of national security and the protection of citizens. These concerns have arisen as Nation States increasingly engage in operations ranging from humanitarian assistance in climate-related disasters to territorial issues exacerbated by changing coastlines, human communities, resource access, and new seaways [Kaye, 2012; Rahman, 2012], 12.6.1). In this regard, increasing sea levels along gently sloping coastlines can have the seemingly perverse outcome that the territorial limits to the maritime jurisdiction of the State might be open to question as the distance from national baselines to the outer limits of the EEZ increases beyond 200 nm over time [Schofield and Arsana, 2012]. Changes in coastal resources may also be coupled with decreasing food security to compound coastal poverty and lead, in some cases, to increased criminal activities such as piracy, IUU fishing, and human, arms and drug trafficking [Kaye, 2012]. While the linkages have not been clearly defined in all cases, it is possible that changes in the Ocean as result of climate change will increase pressure on resources aimed at maintaining maritime security and countering criminal activity, disaster relief operations, and freedom of navigation (12.6.2). National maritime security capacity and infrastructure may also require rethinking as new challenges present themselves as a result of climate change and ocean acidification (12.6.1-2) [Allen and Bergin, 2009; Rahman, 2012]. Opportunities may also arise from changes to international geography such as formation of new ice-free seaways through the Arctic, which may benefit some countries in terms of maintaining maritime security and access (28.2.6). Conversely, such new features may also lead to increasing international tensions as States perceive new vulnerabilities from these changes to geography. Like commercial shipping (30.6.2.3), naval operations in many countries result in significant greenhouse emissions (e.g., the US Navy emits around 2% of the national greenhouse gas emissions, [Mabus, 2010]). As a result, there are a number of programs being implemented by navies around the world to try and reduce their carbon footprint and air pollution such as improving engine efficiency, reducing fouling of vessels, increasing the use of biofuels, and using nuclear technology for power generation, amongst other initiatives. 30.7. Synthesis and Conclusions Evidence that human activities are fundamentally changing the Ocean is virtually certain. Sea temperatures have increased rapidly over the past 60 years at the same time as pH has declined, consistent with the expected influence of rising atmospheric concentrations of CO2 and other greenhouse gases (very high confidence). The rapid rate at which these fundamental physical and chemical parameters of the Ocean are changing is unprecedented within the last 65 Ma (high confidence) and possibly 300 Ma (medium confidence). As the heat content of the Ocean has increased, the Ocean has become more stratified (very likely), although there is considerable regional variability. In some cases, changing surface wind has influenced the extent of mixing and upwelling, although our understanding of where and why these differences occur regionally is uncertain. The changing structure and function of the Ocean has led to changes in parameters such as O2, carbonate ions, and inorganic nutrients concentrations (high confidence). Not surprisingly, these fundamental changes have resulted in responses by key marine organisms, ecosystems and ecological processes, with negative implications for hundreds of millions of people that depend on the ecosystem goods and services provided by the Ocean (very likely). Marine organisms are migrating at rapid rates towards higher latitudes, fisheries are transforming, and many organisms are shifting their reproductive and migratory activity in concert with the changes in temperature and other parameters. Ecosystems such as coral reefs are declining rapidly (high confidence). An extensive discussion of these changes is provided in previous sections and in other chapters of AR5. [INSERT TABLE 30-3 HERE Table 30-3 Key risks to ocean and coastal issues from climate change and the potential for risk reduction through mitigation and adaptation. Key risks are identified based on assessment of the literature and expert judgments made by authors of the various WGII AR5 chapters, with supporting evaluation of evidence and agreement in the referenced chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 are presented for the near-term era of committed climate change (here, for 2030 2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the longer-term era of climate options (here, for 2080 2100), for global mean temperature increases of 2°C and 4°C above pre-industrial levels. For each timeframe, risk levels are estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols. 30.7.1. Key Risks and Vulnerabilities The rapid changes in the physical, chemical, and biological state of the Ocean pose a number of key risks and vulnerabilities for ecosystems, communities, and nations worldwide. Table 30-3 and Figure 30-12 summarize risks and vulnerabilities from climate change and ocean acidification, along with adaptation issues and prospects, and a summary of expert opinion on how these risks will change under further changes in environmental conditions. Rising ocean temperatures are changing the distribution, abundance, and phenology of many marine species and ecosystems, and consequently represent a key risk to food resources, coastal livelihoods, and industries such as tourism and fishing, especially for HLSBS, CBS, STG, and EBUE (Table 30-3, Figure 30-12a-b, 30.5, Box CC-MB, 6.3.1 6.3.4, 7.3.2.4). Key risks involve changes in the distribution and abundance of key fishery species (Figure 30- 12a 2, 4; Figure 30-12b 2; high confidence) as well as the spread of disease and invading organisms, each of which has the potential to impact ecosystems as well as aquaculture and fishing (Table 30-3, 6.3.5, 6.4.1.1, 6.5.3, 7.3.2.4, 7.4.2, 29.5.3, 29.5.4). Adaptation to these changes may be possible in the short-term through dynamic fisheries policy and management (i.e., relocation of fishing effort, Table 30-3), as well as monitoring and responding to potential invading species in coastal settings. The increasing frequency of thermal extremes (Box CC-HS) will also increase the risk that the thermal threshold of corals and other organisms is exceeded on a more frequent basis (especially in CBS, STG, SES, HLSBS, and EUS ocean regions; 30.5, Box CC-CR, 6.2). These changes pose a key risk to vulnerable ecosystems such as mangroves and coral reefs, with potential to have a series of serious impacts on fisheries, tourism, and coastal ecosystem services such as coastal protection (Table 30-3, 30.5, Box CC-CR, 5.4.2.4, 6.3.2. 6.3.5, 6.4.1.3, 7.2.1.2, 29.3.1.2). Genetic adaptation of species to increasing levels of stress may not occur fast enough given fairly long generation times of organisms such as reef building corals and many other invertebrates and fish (Table 30-3). In this case, risks may be reduced by addressing non-climate change related stresses (e.g., pollution, overfishing), although this strategy could have minimal impact if further increases in sea temperature occur (high confidence). The loss of these important coastal ecosystems is associated with the emerging risks associated with the collapse of some coastal fisheries along with livelihoods, food, and regional security (medium confidence). These changes are likely to be exacerbated by other key risks such as coastal inundation and habitat loss due to sea level rise, as well is intensified precipitation events (high confidence) (5.4, Box CC-CR). Adaptation options in this case include engineered coastal defences, re-establishing coastal vegetation such as mangroves, protecting water supplies from salination, and developing strategies for coastal communities to withdraw to less vulnerable locations over time (5.5). The recent decline in O2 concentrations has been ascribed to warming through the effect on ocean mixing and ventilation, as well as the solubility of O2 and its consumption by marine microbes (30.3.2.3, 30.5.7, 6.1.1.3, 6.3.3). This represents a key risk to ocean ecosystems (Figure 30-12a 6, Figure 30-12b 3; medium confidence). These changes increase the vulnerability of marine communities, especially those below the euphotic zone, to hypoxia and ultimately lead to a restriction of suitable habitat (Figure 30-12a 7, high confidence). In the more extreme case, often exacerbated by the contribution of organic carbon from land-based sources, dead zones may form . Decreasing oxygen, consequently, is very likely to increase the vulnerability of fisheries and aquaculture (Figure 30-12b 1, 3; medium confidence), and consequently puts livelihoods at risk, particularly in EBUE (e.g., California and Humboldt Current ecosystems; 30.5.5), SES (e.g., Baltic and Black Seas; 30.5.3) and CBS (e.g., Gulf of Mexico, NE Indian Ocean; 30.5.4, 30.3.2.3). It is very likely that the warming of surface waters has also increased the stratification of the upper ocean by about 4% between 0 and 200 m from 1971 2010 in all oceans north of about 40°S. In many cases, there is significant adaptation opportunity to reduce hypoxia locally by reducing the flow of organic carbon and hence microbial activity within these coastal systems (30.5.4). Relocating fishing effort, and modifying Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 procedures associated with industries like aquaculture, may offer some opportunity to adapt to these changes (likely). Declining O2 concentrations is likely to have significant impacts on DS habitats, where organisms are relatively sensitive to environmental changes of this nature due to the very constant conditions under which they have evolved (30.5.7). Ocean acidification has increased the vulnerability of ocean ecosystems by affecting key aspects of the physiology and ecology of marine organisms (particularly in CBS, STG, and SES regions; Table 30-3; 6.3.2, Box CC-OA). Decreasing pH and carbonate ion concentrations reduce the ability of marine organisms to produce shells and skeletons, and may interfere with a broad range of important processes such as reproduction, navigation, and neural function in a broad range of marine organisms which show minor to major influences of ocean acidification on their biology (30.3.2.2, 6.3.2, Box CC-OA). Natural variability in ocean pH can interact with ocean acidification to create damaging periods of extremes (i.e., high CO2, low O2 and pH), which can have a strong effect on coastal activities such as aquaculture (medium confidence Figure 30-12b 1; Box CC-UP, 6.2). There may be opportunity to adapt aquaculture to increasingly acidic conditions by monitoring natural variability and restricting water intake to periods of optimal conditions. Reducing other non-climate change or ocean acidification associated stresses also represents an opportunity to build greater ecological resilience against the impacts of changing ocean carbonate chemistry. Ocean acidification is also an emerging risk for DS habitats as CO2 continues to penetrate the Ocean, although the impacts and adaptation options are poorly understood and explored. Ocean acidification has heightened importance for some groups of organisms and ecosystems (Box CC-OA). In ecosystems that are heavily dependent on the accumulation of calcium carbonate over time (e.g., coral reefs, Halimeda beds), increasing ocean acidification puts at risk ecosystems services which are critical for hundreds of thousands of marine species, plus people and industries, particularly within CBS, STG and SES (high confidence). Further risks may emerge from the non-linear interaction of different factors (e.g., increasing ocean temperature may amplify effects of ocean acidification, and vice versa) and via the interaction of local stressors with climate change (e.g., interacting changes may lead to greater ecosystems disturbances than each impact on its own). There is an urgent need to understand these types of interactions and impacts, especially given the long time it will take to return ocean ecosystems to pre-industrial pH and carbonate chemistry (i.e., tens of thousands of years, FAQ 30.1, should CO2 emissions continue at the current rate). It is very likely that surface warming has increased stratification of the upper ocean is contributing to the decrease in O2 along with the temperature related decreases in oxygen solubility (WGI 3.8.3). Changes to wind speed, wave height, and storm intensity influence the location and rate of mixing within the upper layers of the Ocean and hence the concentration of inorganic nutrients (e.g., in EBUE, EUS; Figure 30-12a 1, 3). These changes to ocean structure increase the risks and vulnerability of food webs within the Ocean. However, our understanding of how primary productivity is going to change in a warming and more acidified ocean is limited, as is our understanding of how upwelling will respond to changing surface wind as the world continues to warm (Box CC-PP, Box CC-UP). As already discussed, these types of changes can have implications for the supply of O2 into the Ocean and the upward transport of inorganic nutrients to the euphotic zone. While our understanding is limited, there is significant potential for regional increases in wind speed to result in greater rates of upwelling and the supply of inorganic nutrients to the photic zone. While this may increase productivity of phytoplankton communities and associated fisheries, greater rates of upwelling can increase the risk of hypoxic conditions developing at depth as excess primary production sinks into the Ocean and stimulates microbial activity at depth (Table 30-3, 30.3.2.3, 30.5.5, 6.1.1.3). Changes in storm intensity can increase the risk of damage to shipping and industrial infrastructure, which increases the risk of accidents and delays to the transport of products between countries, security operations, and the extraction of minerals from coastal and oceanic areas (30.6.2, [IPCC, 2012]. The proliferation of key risks and vulnerabilities to the goods and services provided by ocean ecosystems as a result of ocean warming and acidification generate a number of key risks for the citizens of almost every nation. Risks to food security and livelihoods are expected to increase over time, aggravating poverty and inequity (Table 30-3). As these problems increase, regional security is likely to deteriorate as disputes over resources increase, along with increasing insecurity of food and nutrition (Table 30-3, [IPCC, 2012], 30.6.5, 12.4-12.6, 29.3). Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 30.7.2. Global Frameworks for Decision-Making Global frameworks for decision-making are central to management of vulnerability and risk at the scale and complexity of the world s oceans. General frameworks and conventions for policy development and decision- making within oceanic and coastal regions are important in terms of the management of stressors not directly due to ocean warming or acidification, but which may influence the outcome of these two factors. Table 30-3 and Table 30-4 outlines a further set of challenges arising from multiple interacting stressors, as well as potential risks and vulnerabilities, ramifications, and adaptation options. In the latter case, examples of potential global frameworks and initiatives for initiating and managing these adaptation options are described. These frameworks represent opportunities for global cooperation and the development of international, regional, and national policy responses to the challenges posed by the changing ocean [Kenchington and Warner, 2012; Tsamenyi and Hanich, 2012; Warner and Schofield, 2012]. [INSERT TABLE 30-4 HERE Table 30-4: Ramifications, adaptation options and frameworks for decision-making for ocean regions. Symbols are as follows: T = sea temperature; UW = upwelling; OA = ocean acidification; NU = nutrient concentration; IC = ice cover; SS = storm strength; SLR = sea level rise ( = Increased; = decreased; italics = uncertain). Acronyms are: CBD (Convention on Biological Diversity); CTI (Coral Triangle Initiative); GEF (Global Environment Facility); IHO (International Hydrographic Organization); ILO (International Labor Organization); IOM (International Organization of Migration); ISPS (International Ship and Port Facility Security); MARPOL (International Convention for the Prevention of Pollution From Ships); PACC (Pacific Adaptation to Climate Change Project); PEMSEA (Partnerships in Environmental Management for the Seas of East Asia); RFMO (Regional Fisheries Management Organizations); SPREP (Secretariat of the Pacific Regional Environment Programme); UNCLOS (United Nations Convention on the Law of the Sea); UNHCR (United Nations High Commissioner for Refugees); UNSFSA (Straddling Fish Stocks Agreement); and WHO (World Health Organization).] The United Nations Convention on the Law of the Sea (UNCLOS) was a major outcome of the third UN Conference on the Law of the Sea (UNCLOS III). The European Union and 164 countries have joined in the Convention. UNCLOS replaced earlier frameworks that were built around the freedom of the seas concept and which limited territorial rights to 3 nm off a coastline. UNCLOS provides a comprehensive framework for the legitimate use of the Ocean and its resources, including maritime zones, navigational rights, protection and preservation of the marine environment, fishing activities, marine scientific research, and mineral resource extraction from the seabed beyond national jurisdiction. The relationship between climate change and UNCLOS is not clear and depends on interpretation of the key elements within the UNFCCC (United Nations Framework Convention for Climate Change) and Kyoto Protocol [Boyle, 2012]. However, UNCLOS provides mechanisms to help structural adaptation in response to challenges posed by climate change. In a similar way, there is a wide range of other policy and legal frameworks that structure and enable responses to the outcomes of rapid anthropogenic climate change in the Ocean. There are many existing international conventions and agreements that explicitly recognize climate change (Table 30-4). The United Nations Straddling Fish Stocks Agreement (UNSFSA) aims at enhancing international cooperation of fisheries resources, with an explicit understanding under article 6 that management needs to take account existing and predicted oceanic, environmental and socio-economic conditions and to undertake relevant research, including surveys of abundance, biomass surveys, hydro-acoustic surveys, research on environmental factors affecting stock abundance, and oceanographic and ecological studies (Annex one, article 3). International conventions such as these will become increasingly important as changes to the distribution and abundance of fisheries are modified by climate change and ocean acidification. Global frameworks for decision-making are increasingly important in the case of the Ocean, most of which falls outside national boundaries [Elferink, 2012; Warner, 2012]. Approximately 64% of the Ocean (40% of the Earth s surface) is outside EEZs and continental shelves of the world's nations (high seas and seabed beyond national jurisdiction). With rapidly increasing levels of exploitation, there are increasing calls for more effective decision frameworks aimed at regulating fishing and other activities (e.g., bio-prospecting) within these ocean commons . These international frameworks will become increasingly valuable as nations respond to impacts on fisheries resources that stretch across national boundaries. One such example is the multilateral cooperation that was driven Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 by President Yudhoyono of Indonesia in August 2007 and led to the Coral Triangle Initiative on Coral Reefs, Fisheries, and Food Security (CTI), which involves region-wide (involving 6.8 million km2 including 132,800 km of coastline) cooperation between the governments of Indonesia, the Philippines, Malaysia, Papua New Guinea, the Solomon Islands, and Timor Leste on reversing the decline in coastal ecosystems such as coral reefs [Clifton, 2009; Hoegh-Guldberg et al., 2009; Veron et al., 2009]. Partnerships, such as the CTI, have the potential to provide key frameworks to address issues such as interaction between the over-exploitation of coastal fishing resources and the recovery of reefs from mass coral bleaching and mortality, and the implications of the movement of valuable fishery stocks beyond waters under national jurisdiction. An initiative called the Global Partnership for Oceans set out (www.globalpartnershipforoceans.org, March 28, 2012) to establish a global framework with which to share experience, resources and expertise, as well as to engage governments, industry, civil, and public sector interests in both understanding and finding solutions to key issues such as overfishing, pollution, and habitat destruction [GPO, 2012]. Similarly, the Areas Beyond National Jurisdiction (ABNJ, Global Environment Facility) Initiative has been established to promote the efficient, collaborative, and sustainable management of fisheries resources and biodiversity conservation across the Ocean. Global partnerships are also essential for providing support to the many nations that often do not have the scientific or financial resources to solve the challenges that lie ahead [Busby, 2009; Mertz et al., 2009]. In this regard, international networks and partnerships are particularly significant in terms of assisting nations in developing local adaptation solutions to their ocean resources. By sharing common experiences and strategies through global networks, nations have the chance to tap into a vast array of options with respect to responding to the negative consequences of climate change and ocean acidification on the world s ocean and coastal resources. 30.7.3. Emerging Issues, Data Gaps, and Research Needs While there has been an increase in the number of studies being undertaken to understand the physical, chemical, and biological changes within the Ocean in response to climate change and ocean acidification, the number of marine studies still lag significantly behind terrestrial and atmospheric studies [Hoegh-Guldberg and Bruno, 2010; Poloczanska et al., 2013]. Rectifying this gap should be a major international objective given the importance of the Ocean in terms of understanding and responding to future changes and consequences from ocean warming and acidification. 30.7.3.1. Changing Variability and Marine Impacts Understanding the long-term variability of the Ocean is critically important in terms of the detection and attribution of changes to climate change (30.3, 30.5.8), but also in terms of the interaction between variability and anthropogenic climate change. Developing instrument systems that expand the spatial and temporal coverage of the Ocean and key processes will be critical to documenting and understanding its behavior under further increases in average global temperature and changes the atmospheric concentration of CO2. International collaborations such as the Argo network of oceanographic floats are rapidly improving our understanding of the physical behavior of the Ocean and will provide important insight into its long-term subsurface variability [Schofield et al., 2013]. 30.7.3.2. Surface Wind, Storms, and Upwelling Improving our understanding of the potential behavior of surface wind in a warming world is needed for improving our understanding of how upwelling will change in key regions (e.g., EUS, EBUE; Box CC-UP). Understanding these changes will provide important information for future fisheries management but will also illuminate the potential risks of intensified upwelling leading to hypoxia at depth and the potential expansion of dead zones (30.3.2; 30.5.2-30.5.4). Understanding surface wind in a warming climate will also yield important information on surface mixing as well as how surface wave height might also vary, improving our understanding of potential interactions in coastal areas between wind, waves, and sea level rise (30.3.1). Given the importance of mixing and upwelling to the supply of inorganic nutrients to the surface layers of the ocean, understanding these important Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 phenomena at the ocean-atmosphere interface will provide important insight into how ocean warming and acidification are likely to impact ecosystems, food webs and ultimately central important fisheries such as those found along the west coasts of Africa and the Americas. 30.7.3.3. Declining O2 Concentrations The declining level of O2 in the Ocean is an emerging issue of major importance (30.3.2). Developing a better understanding of the role and temperature sensitivity of microbial systems in determining O2 concentrations will enable a more coherent understanding of the changes and potential risks to marine ecosystems. Given the importance of microbial systems to the physical, chemical, and biological characteristics of the Ocean, it is extremely important that these systems receive greater focus, especially with regards to their response to ocean warming and acidification. This is particularly important for the DS (>1000 m), which is the most extensive habitat on the planet. In this respect, increasing our understanding of DS habitats and how they may be changing under the influence of climate change and ocean acidification is of great importance. Linkages between changes occurring in the surface layers and those associated with the DS are particularly important in light of our need to understand how rapidly changes are occurring and what the implications are for the metabolic activity and O2 content of DS habitats. 30.7.3.4. Ocean Acidification The rapid and largely unprecedented changes to ocean acidification represent an emerging issue given the central importance of pH and the concentration of ions such as carbonate in the biology of marine organisms (Box CC-OA). Despite the relatively short history of research on this issue, there are already a large number of laboratories and field studies that demonstrate a large range of effects across organisms, processes, and ecosystems. Key gaps [Gattuso et al., 2011] remain in our understanding of how ocean acidification will interact with other changes in the Ocean, and whether or not biological responses to ocean acidification are necessarily linear. The vulnerability of fishery species (e.g., mollusks) to ocean acidification represents an emerging issue, with a need for research to understand and develop strategies for industry to minimize the impacts. Understanding of how carbonate structures like coral reefs and Halimeda beds, will respond to a rapidly acidifying ocean represents a key gap and research need, especially in understanding the rate at which consolidated carbonate structures and related habitats are likely to erode and dissolve. Interactions between ocean acidification, upwelling, and decreasing O2 represent additional areas of concern and research. There is also a need to improve our understanding of the socio-economic ramifications of ocean acidification [Turley et al., 2011; Hilmi et al., 2013]. 30.7.3.5. Net Primary Productivity Oceanic phytoplankton are responsible for 50% of global net primary productivity. However, our understanding of how oceanic primary production is likely to change in a warmer and more acidified ocean is uncertain (Box CC-PP; Box CC-UP). Changes in net primary productivity will resonate through food webs and ultimately affect fish production. Given the central role that primary producers and their associated ecological processes play in ocean ecosystem functioning, it is crucial that we improve our understanding of how net primary productivity is likely to vary at global and regional levels (30.5.2, 30.5.5). At the same time, understanding how plankton communities will vary spatially and temporarily will be important in any attempt to understand how fish populations will fare in a warmer and more acidified ocean. The research challenge is to determine when and where net primary production is expected to change, coupled with research on adaptation strategies for changes to the global distribution of seafood procurement, management and food security. 30.7.3.6. Movement of Marine Organisms and Ecosystems Marine organisms are moving, generally towards higher latitudes or deeper waters, consistent with the expectation of a warming ocean. Our current understanding of which organisms and ecosystems are moving, and the Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 ramifications for reorganisation of ecosystems and communities, and the implications for nations is uncertain at best. Given the implications for fisheries, invasive species, and the spread of disease, it is an imperative that our understanding of the movement of ecosystems is improved. Documentation of species responses and a deeper understanding of the processes that lead to persistent range shifts, and a focus on the ecosystem, social and economic implications of range shifts is a research need. 30.7.3.7. Understanding Cumulative and Synergistic Impacts Understanding cumulative and synergistic impacts is poorly developed for ocean systems. Much of our understanding has been built on experimental approaches that are focused on single stressors that respond gradually without interaction or have impacts that accumulate over time (Table 30-3). Multifactorial experiments exploring the impact of combined variables (e.g., elevated temperature and acidification at the same time) will enable more realistic projections of the future to be established. Equally, developing a better understanding of how biological and ecological responses change in relation to key environmental variables should also be a goal of future research. In this regard, assumptions that responses are likely to be gradual and linear over time ultimately have little basis, yet are widespread within the scientific literature. 30.7.3.8. Reorganization of Ecosystems and Food Webs The pervasive influence of ocean warming and acidification on the distribution, abundance, and function of organisms and processes has and will continue to drive the reorganization of ecosystems and food webs (virtually certain) ([Hoegh-Guldberg and Bruno, 2010; Poloczanska et al., 2013], Box CC-MB). One of the inevitable outcomes of differing tolerances and responses to climate change and ocean acidification is the development of novel assemblages of organisms in the near future. Such communities are likely to have no past or contemporary counterparts, and will consequently require new strategies for managing coastal areas and fisheries. Changes to a wide array of factors related or not related to climate change have the potential to drive extremely complex changes in community structure and, consequently, food web dynamics. Developing a greater capability for detecting and understanding these changes will be critical for future management of ocean and coastal resources. 30.7.3.9. Socio-ecological Resilience Many communities depend on marine ecosystems for food and income yet our understanding of the consequences of environmental degradation is poor. For example, while there is high confidence that coral reefs will continue to deteriorate at current rates of climate change and ocean acidification [Gardner et al., 2003; Bruno and Selig, 2007; De ath et al., 2012], there is relatively poor understanding of the implications for the hundreds of millions of people who depend on these important coastal ecosystems for food and livelihoods. Improving our understanding of how to reinforce socio-ecological resilience in communities affected by the deterioration of key coastal and oceanic ecosystems is central to developing effective adaptation responses to these growing challenges (30.6 Table 30-3, Table 30-4). Frequently Asked Questions FAQ 30.1: Can we reverse the climate change impacts on the ocean? [to be inserted after Section 30.3.2] In less than 150 years, greenhouse gas emissions have resulted in such major physical and chemical changes in our oceans that it will take thousands of years to reverse them. There are a number of reasons for this. Given its large mass and high heat capacity, the ability of the Ocean to absorb heat is 1000 times larger than that of the atmosphere. The Ocean has absorbed at least nine tenths of the Earth s heat gain between 1971 and 2010. To reverse that heating, the warmer upper layers of the Ocean have to mix with the colder deeper layers. That mixing can take up to 1000 years. This means it will take centuries to millennia for deep ocean temperatures to warm in response to today s surface conditions, and at least as long for ocean warming to reverse after atmospheric greenhouse gas Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 concentrations decrease (virtually certain). But climate change-caused alteration of basic conditions in the Ocean is not just about temperature. The Ocean becomes more acidic as more CO2 enters it and will take tens of thousands of years to reverse these profound changes to the carbonate chemistry of the ocean (virtually certain). These enormous physical and chemical changes are producing sweeping and profound changes in marine ecosystems. Large and abrupt changes to these ecosystems are unlikely to be reversible in the short to medium term (high confidence). FAQ30.2: Does slower warming mean less impact on plants and animals? [to be inserted in Section 30.4] The greater thermal inertia of the Ocean means that temperature anomalies and extremes are lower than those seen on land. This does not necessarily mean that impacts of ocean warming are less for the ocean than for land. A large body of evidence reveals that small amounts of warming in the Ocean can have large effects on ocean ecosystems. For example, relatively small increases in sea temperature (as little as 1 2°C) can cause mass coral bleaching and mortality across hundreds of square kilometers of coral reef (high confidence). Other analyses have revealed that increased temperatures are spreading rapidly across the world s oceans (measured as the movement of bands of equal water temperature or isotherms). This rate of warming presents challenges to organisms and ecosystems as they try to migrate to cooler regions as the Ocean continues to warm. Rapid environmental change also poses steep challenges to evolutionary processes, especially where long-lived organisms such as corals and fish are concerned (high confidence). FAQ30.3: How will marine primary productivity change? [to be inserted after Section 30.5.2.2] Drifting microscopic plants known as phytoplankton are the dominant marine primary producers, at the base of the marine food chain. Their photosynthetic activity is critically important to life in general. It provides oxygen, supports marine food webs, and influences global biogeochemical cycles. Changes in marine primary productivity in response to climate change remain the single biggest uncertainty in predicting the magnitude and direction of future changes in fisheries and marine ecosystems (low confidence). Changes have been reported to a range of different ocean systems (e.g., High Latitude Spring Bloom Systems, Sub-tropical Gyre Systems, Equatorial Upwelling Systems, and Eastern Boundary Upwelling Ecosystems), some of which are consistent with changes in ocean temperature, mixing, and circulation. However, direct attribution of these changes to climate change is made difficult by long-term patterns of variability that influence productivity of different parts of the Ocean (e.g., Pacific Decadal Oscillation). Given the importance of this question for ocean ecosystems and fisheries, longer time series studies to understand how these systems are changing as a result of climate change are a priority (high agreement). FAQ30.4: Will climate change cause dead zones in the oceans? [to be inserted after Section 30.5.5.2] Dissolved oxygen is a major determinant of the distribution and abundance of marine organisms. Dead zones are persistent hypoxic conditions where the water doesn t have enough dissolved oxygen to support oxygen-dependent marine species. These areas exist all over the world and are expanding, with impacts on coastal ecosystems and fisheries (high confidence). Dead zones are caused by several factors, particularly eutrophication where too many nutrients run off coastal cities and agricultural areas into rivers that carry these materials out to sea. This stimulates primary production leading to a greater supply of organic carbon, which can sink into the deeper layers of the ocean. As microbial activity is stimulated, there is a sharp reduction in dissolved oxygen levels and an increased risk of dead zones (high confidence). Climate change can influence the distribution of dead zones by increasing water temperature and hence microbial activity, as well as reducing mixing of the ocean (i.e., increasing layering or stratification) of the Ocean which have different temperatures, densities, salinities and reducing mixing of oxygen-rich surface layers into the deeper parts of the Ocean. In other areas, increased upwelling can lead to stimulated productivity, which can also lead to more organic carbon entering the deep ocean, where it is consumed, decreasing oxygen levels (medium confidence). Managing local factors such as the input of nutrients into coastal regions can play an important role in reducing the rate at which dead zones are spreading across the world s oceans (high agreement). FAQ30.5: How can we use non-climate factors to manage climate change impacts on the oceans? [to be inserted after Section 30.7.1] Like most natural system, the Ocean is exposed to a range of stresses that may or may not be related to climate change. Human activities can result in pollution, eutrophication (too many nutrients), habitat destruction, invasive species, destructive fishing, and over-exploitation of marine resources. Sometimes, these activities can increase the impacts of climate change, although they can, in a few circumstances, dampen the effects as well. Understanding Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 how these factors interact with climate change and ocean acidification is important in its own right. However, reducing the impact of these non-climate factors may reduce the overall rate of change within ocean ecosystems. Building ecological resilience through ecosystem-based approaches to the management of the marine environment, for example, may pay dividends in terms of reducing and delaying the effects of climate change (high confidence). Cross-Chapter Boxes Box CC-CR. Coral Reefs [Jean-Pierre Gattuso (France), Ove Hoegh-Guldberg (Australia), Hans-Otto Pörtner (Germany)] Coral reefs are shallow-water ecosystems that consist of reefs made of calcium carbonate which is mostly secreted by reef-building corals and encrusting macroalgae. They occupy less than 0.1% of the ocean floor yet play multiple important roles throughout the tropics, housing high levels of biological diversity as well as providing key ecosystem goods and services such as habitat for fisheries, coastal protection and appealing environments for tourism (Wild et al., 2011). About 275 million people live within 30 km of a coral reef (Burke et al., 2011) and derive some benefits from the ecosystem services that coral reefs provide (Hoegh-Guldberg, 2011) including provisioning (food, livelihoods, construction material, medicine), regulating (shoreline protection, water quality), supporting (primary production, nutrient cycling) and cultural (religion, tourism) services. This is especially true for the many coastal and small island nations in the world s tropical regions (29.3.3.1). Coral reefs are one of the most vulnerable marine ecosystems (high confidence; 5.4.2.4, 6.3.1, 6.3.2, 6.3.5, 25.6.2, and 30.5) and more than half of the world s reefs are under medium or high risk of degradation (Burke et al., 2011). Most human-induced disturbances to coral reefs were local until the early 1980s (e.g., unsustainable coastal development, pollution, nutrient enrichment and overfishing) when disturbances from ocean warming (principally mass coral bleaching and mortality) began to become widespread (Glynn, 1984). Concern about the impact of ocean acidification on coral reefs developed over the same period, primarily over the implications of ocean acidification for the building and maintenance of the calcium carbonate reef framework (Box CC-OA). [INSERT FIGURE CR-1 HERE Figure CR-1: A and B: the same coral community before and after a bleaching event in February 2002 at 5 m depth, Halfway Island, Great Barrier Reef. Coral cover at the time of bleaching was 95% bleached almost all of it severely bleached, resulting in mortality of 20.9% (Elvidge et al., 2004). Mortality was comparatively low due in part because these coral communities were able to shuffle their symbiont to more thermo-tolerant types (Berkelmans and van Oppen, 2006; Jones et al., 2008). C and D: three CO2 seeps in Milne Bay Province, Papua New Guinea show that prolonged exposure to high CO2 is related to fundamental changes in the ecology of coral reefs (Fabricius et al., 2011), including reduced coral diversity (-39%), severely reduced structural complexity (-67%), lower density of young corals (-66%) and fewer crustose coralline algae (-85%). At high CO2 sites (panel D; median pHT ~7.8), reefs are dominated by massive corals while corals with high morphological complexity are underrepresented compared with control sites (D; median pH ~8.0). Reef development ceases at pHT values below 7.7. pHT: pH on the total scale. E: temporal trend in coral cover for the whole Great Barrier Reef over the period 1985 2012 (N, number of reefs, mean +/- 2 standard errors; De'ath et al., 2012). F: composite bars indicate the estimated mean coral mortality for each year, and the sub-bars indicate the relative mortality due to crown-of-thorns starfish, cyclones, and bleaching for the whole Great Barrier Reef (De'ath et al., 2012). Photo credit: R. Berkelmans (A and B) and K. Fabricius (C and D).] A wide range of climatic and non-climatic drivers affect corals and coral reefs and negative impacts have already been observed (5.4.2.4, 6.3.1, 6.3.2, 25.6.2.1, 30.5.3, 30.5.6). Bleaching involves the breakdown and loss of endosymbiotic algae, which live in the coral tissues and play a key role in supplying the coral host with energy (see 6.3.1. for physiological details and 30.5 for a regional analysis). Mass coral bleaching and mortality, triggered by positive temperature anomalies (high confidence), is the most widespread and conspicuous impact of climate change (Figure CR-1A and B, Figure 5-3; 5.4.2.4, 6.3.1, 6.3.5, 25.6.2.1, 30.5 and 30.8.2). For example, the level of thermal stress at most of the 47 reef sites where bleaching occurred during 1997-98 was unmatched in the period 1903 to 1999 (Lough, 2000). Ocean acidification reduces biodiversity (Figure CR-1C and D) and the calcification rate of Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 corals (high confidence; 5.4.2.4, 6.3.2, 6.3.5) while at the same time increasing the rate of dissolution of the reef framework (medium confidence; 5.2.2.4) through stimulation of biological erosion and chemical dissolution. Taken together, these changes will tip the calcium carbonate balance of coral reefs towards net dissolution (medium confidence; 5.4.2.4). Ocean warming and acidification have synergistic effects in several reef-builders (5.2.4.2, 6.3.5). Taken together, these changes will erode habitats for reef-based fisheries, increase the exposure of coastlines to waves and storms, as well as degrading environmental features important to industries such as tourism (high confidence; 6.4.1.3, 25.6.2, 30.5). A growing number of studies have reported regional scale changes in coral calcification and mortality that are consistent with the scale and impact of ocean warming and acidification when compared to local factors such as declining water quality and overfishing (Hoegh-Guldberg et al., 2007). The abundance of reef building corals is in rapid decline in many Pacific and SE Asian regions (very high confidence, 1-2% per year for 1968-2004; Bruno and Selig, 2007). Similarly, the abundance of reef-building corals has decreased by over 80% on many Caribbean reefs (1977 to 2001; Gardner et al., 2003), with a dramatic phase shift from corals to seaweeds occurring on Jamaican reefs (Hughes, 1994). Tropical cyclones, coral predators and thermal stress-related coral bleaching and mortality have led to a decline in coral cover on the Great Barrier Reef by about 51% between 1985 and 2012 (Figure CR-1E and F). Although less well documented, benthic invertebrates other than corals are also at risk (Przeslawski et al., 2008). Fish biodiversity is threatened by the permanent degradation of coral reefs, including in a marine reserve (Jones et al., 2004). Future impacts of climate-related drivers (ocean warming, acidification, sea level rise as well as more intense tropical cyclones and rainfall events) will exacerbate the impacts of non-climate related drivers (high confidence). Even under optimistic assumptions regarding corals being able to rapidly adapt to thermal stress, one-third (9 to 60%, 68% uncertainty range) of the world s coral reefs are projected to be subject to long-term degradation (next few decades) under the RCP3-PD scenario (Frieler et al., 2013). Under the RCP4.5 scenario, this fraction increases to two-thirds (30 to 88%, 68% uncertainty range). If present day corals have residual capacity to acclimate and/or adapt, half of the coral reefs may avoid high frequency bleaching through 2100 (limited evidence, limited agreement; Logan et al., 2013). Evidence of corals adapting rapidly, however, to climate change is missing or equivocal (Hoegh-Guldberg, 2012). Damage to coral reefs has implications for several key regional services: Resources: Coral reefs account for 10 to 12% of the fish caught in tropical countries, and 20 to 25% of the fish caught by developing nations (Garcia and Moreno, 2003). Over half (55%) of the 49 island countries considered by Newton et al. (2007) are already exploiting their coral reef fisheries in an unsustainable way and the production of coral reef fish in the Pacific is projected to decrease 20% by 2050 under the SRES A2 emissions scenario (Bell et al., 2013). Coastal protection: Coral reefs contribute to protecting the shoreline from the destructive action of storm surges and cyclones (Sheppard et al., 2005), sheltering the only habitable land for several island nations, habitats suitable for the establishment and maintenance of mangroves and wetlands, as well as areas for recreational activities. This role is threatened by future sea level rise, the decrease in coral cover, reduced rates of calcification and higher rates of dissolution and bioerosion due to ocean warming and acidification (5.4.2.4, 6.4.1, 30.5). Tourism: More than 100 countries benefit from the recreational value provided by their coral reefs (Burke et al., 2011). For example, the Great Barrier Reef Marine Park attracts about 1.9 million visits each year and generates A$ 5.4 billion to the Australian economy and 54,000 jobs (90% in the tourism sector; Biggs, 2011). Coral reefs make a modest contribution to the Global Domestic Product but their economic importance can be high at the country and regional scales (Pratchett et al., 2008). For example, tourism and fisheries represent 5% of the GDP of South Pacific islands (average for 2001-2011; Laurans et al., 2013). At the local scale, these two services provided in 2009-2011 at least 25% of the annual income of villages in Vanuatu and Fiji (Pascal, 2011; Laurans et al., 2013). Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Isolated reefs can recover from major disturbance, and the benefits of their isolation from chronic anthropogenic pressures can outweigh the costs of limited connectivity (Gilmour et al., 2013). Marine protected areas (MPAs) and fisheries management have the potential to increase ecosystem resilience and increase the recovery of coral reefs after climate change impacts such as mass coral bleaching (McLeod et al., 2009). Although they are key conservation and management tools, they are unable to protect corals directly from thermal stress (Selig et al., 2012) suggesting that they need to be complemented with additional and alternative strategies (Rau et al., 2012; Billé et al., 2013). While MPA networks are a critical management tool, they should be established considering other forms of resource management (e.g., fishery catch limits and gear restrictions) and integrated ocean and coastal management to control land-based threats such as pollution and sedimentation. There is medium confidence that networks of highly protected areas nested within a broader management framework can contribute to preserving coral reefs under increasing human pressure at local and global scales (Salm et al. 2006). Locally, controlling the input of nutrients and sediment from land is an important complementary management strategy (Mcleod et al., 2009) because nutrient enrichment can increase the susceptibility of corals to bleaching (Wiedenmann et al., 2012) and coastal pollutants enriched with fertilizers can increase acidification (Kelly et al., 2011). In the long term, limiting the amount of ocean warming and acidification is central to ensuring the viability of coral reefs and dependent communities (high confidence; 5.2.4.4, 30.5). CC-CR References Bell J. D., A. Ganachaud, P.C. Gehrke, S.P. Griffiths, A.J. Hobdaym O. Hoegh-Guldberg, J.E. Johnson, R. Le Borgne, P. Lehodey, J.M. Lough, R.J. Matear, T.D. Pickering, M.S. Pratchett, A. Sen Gupta, I. Senina I. and M. Waycott . 2013: Mixed responses of tropical Pacific fisheries and aquaculture to climate change. Nature Climate Change 3, 591-591. Berkelmans R. and M.J.H. van Oppen, 2006: The role of zooxanthellae in the thermal tolerance of corals: a nugget of hope for coral reefs in an era of climate change. In: Proceedings of the Royal Society of London. Series B: Biological Sciences, 273, 2305-2312. Biggs D., 2011. Case study: the resilience of the nature-based tourism system on Australia s Great Barrier Reef. Report prepared for the Australian Department of Sustainability, Environment, Water, Population and Communities on behalf of the State of the Environment 2011 Committee Government. 32 p. Billé R., R. Kelly, E. Harrould-Kolieb, D. Herr, F. Joos, K.J. Kroeker, D. Laffoley, A. Oschlies and J.P. Gattuso , 2013: Taking action against ocean acidification: a review of management and policy options. Environmental Management, 52, 761-779. Bruno J. F. and E.R. Selig, 2007: Regional decline of coral cover in the Indo-Pacific: timing, extent, and subregional comparisons. PLoS ONE, 2(8), e711. Burke L. M., K. Reytar, M. Spalding and A. Perry, 2011: Reefs at risk revisited. World Resources Institute, Washington, DC:. p.114. De ath G., K.E. Fabricius, H. Sweatman and M. Puotinen, 2012: The 27-year decline of coral cover on the Great Barrier Reef and its causes. In: Proceedings of the National Academy of Science U.S.A. 109, 17995-17999. Elvidge C., J. Dietz, R. Berkelmans, S. Andréfouët, W. Skirving, A. Strong and B. Tuttle, 2004: Satellite observation of Keppel Islands (Great Barrier Reef) 2002 coral bleaching using IKONOS data. Coral Reefs 23, 123-132. Fabricius K. E., C. Langdon, S. Uthicke, C. Humphrey, S. Noonan, G. De ath, R. Okazak, N. Muehllehner, M.S. Glas and J.M. Lough, 2011: Losers and winners in coral reefs acclimatized to elevated carbon dioxide concentrations. Nature Climate Change 1, 165-169. Frieler K., M. Meinshausen, A. Golly, M. Mengel, K. Lebek, S.D. Donner and O. Hoegh-Guldberg, 2013: Limiting global warming to 2 °C is unlikely to save most coral reefs. Nature Climate Change 3, 165-170. Garcia S. M. and I. de Leiva Moreno, 2003: Global overview of marine fisheries. In: Responsible fisheries in the marine ecosystem, [Sinclair M. and Valdimarsson G. (eds.)], Wallingford: CABI pp. 1-24. Gardner T. A., I.M. Cote, J.A. Gill, A. Grant and A.R. Watkinson, 2003: Long-term region-wide declines in Caribbean corals. Science 301(5635), 958-960. Gilmour J. P., LD. Smith, A.J. Heyward, A.H. Baird and M.S. Pratchett M. S., 2013: Recovery of an isolated coral reef system following severe disturbance. Science 340, 69-71. Glynn P. W., 1984: Widespread coral mortality and the 1982-83 El Nino warming event. Environmental Conservation 11, 133-146. Hoegh-Guldberg O., 2011: Coral reef ecosystems and anthropogenic climate change. Regional Environmental Change 11, 215-227. Hoegh-Guldberg O., 2012: The adaptation of coral reefs to climate change: is the Red Queen being outpaced? Scientia Marina 76, 403-408. Hoegh-Guldberg, O., P. J. Mumby, A. J. Hooten, R. S. Steneck, P. Greenfield, E. Gomez, C. D. Harvell, P. F. Sale, A. J. Edwards, K. Caldeira, N. Knowlton, C. M. Eakin, R. Iglesias-Prieto, N. Muthiga, R. H. Bradbury, A. Dubi, and M. E. Hatziolos, 2007: Coral reefs under rapid climate change and ocean acidification. Science 318, 1737-1742. Hughes T. P., 1994. Catastrophes, phase shifts, and large-scale degradation of a Caribbean coral reef. Science 265(5178), 1547-1551. Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Jones A. M., R. Berkelmans, M.J. van Oppen, J.C. Mieog and W. Sinclair, 2008: A community change in the algal endosymbionts of a scleractinian coral following a natural bleaching event: field evidence of acclimatization. In: Proceedings of the Royal Society of London. Series B: Biological Sciences 275, 1359-1365. Jones G. P., M.I. McCormick, M. Srinivasan and J.V. Eagle, 2004: Coral decline threatens fish biodiversity in marine reserves. In: Proceedings of the National Academy of Science U.S.A. 101, 8251-8253. Kelly R. P., M.M. Foley, W.S. Fisher, R.A. Feely, B.S. Halpern, G.G. Waldbusser and M.R. Caldwell, 2011: Mitigating local causes of ocean acidification with existing laws. Science 332, 1036-1037. Laurans Y., N. Pascal, T. Binet, L. Brander, E. Clua, G. David, D. Rojat and A. Seidl, 2013: Economic valuation of ecosystem services from coral reefs in the South Pacific: taking stock of recent experience. Journal of Environmental Management 116C, 135-144. Logan C. A., J.P. Dunne, C.M. Eakin and S.D. Donner, 2013: Incorporating adaptation and acclimatization into future projections of coral bleaching. Global Change Biology, doi:10.1111/gcb.12390. Lough J. M., 2000: 1997-98: Unprecedented thermal stress to coral reefs? Geophysical Research Letters. 27(23), 3901-3904. McLeod E., R. Salm, A. Green and J. Almany, 2009: Designing marine protected area networks to address the impacts of climate change. Frontiers in Ecology and the Environment 7, 362-370. Newton K., I. M. Côté, G.M. Pilling G, S. Jennings and N.K. Dulvy, 2007: Current and future sustainability of island coral reef fisheries. Current Biology 17. 655-658. Pascal N., 2011. Cost-benefit analysis of community-based marine protected areas: 5 case studies in Vanuatu. Moorea, French Polynesia: CRISP-CRIOBE, 107p. Pratchett M. S., P.L. Munday and S.K. Wilson, 2008: Effects of climate-induced coral bleaching on coral-reef fishes- Ecological and economic consequences. Oceanography and Marine Biology: an Annual Review 46, 251-296. Przeslawski R., A. Ahyong, M. Byrne, G. Worheide and P. Hutchings, 2008: Beyond corals and fish: the effects of climate change on noncoral benthic invertebrates of tropical reefs. Global Change Biology 14, 2773-2795. Rau G. H., E.L. McLeod and O. Hoegh-Guldberg, 2012: The need for new ocean conservation strategies in a high-carbon dioxide world. Nature Climate Change 2, 720-724. Salm RV, T. Done and E. Mcleod, 2006: Marine protected area planning in a changing climate. In: Coral Reefs and Climate Change: Science and Management. [Phinney, J.T., Hoegh- Guldberg O, J. Kleypas, et al. (eds)].Washington, DC: American Geophysical Union 244 pp.. Selig E. R., K.S. Casey and J.F. Bruno, 2012: Temperature-driven coral decline: the role of marine protected areas. Global Change Biology 18, 1561-1570. Sheppard C., D.J. Dixon, M. Gourlay, A. Sheppard and R. Payet, 2005: Coral mortality increases wave energy reaching shores protected by reef flats: examples from the Seychelles. Estuarine, Coastal and Shelf Science 64, 223-234. Wiedenmann J., C. D Angelo, E.G. Smith, A.N. Hunt, F.E. Legiret, A.D. Postle and E.P. Achterberg, 2013: Nutrient enrichment can increase the susceptibility of reef corals to bleaching. Nature Climate Change 3, 160-164. Wild C., O. Hoegh-Guldberg, M.S. Naumann, M. Florencia Colombo-Pallotta, M. Ateweberhan, W.K. Fitt , R. Iglesias-Prieto, C. Palmer, J.C. Bythell, J.-C.Ortiz, Y. Loya and R. van Woesik, 2011: Climate change impedes scleractinian corals as primary reef ecosystem engineers. Marine and Freshwater Research 62, 205-215. Box CC-HS. Heat Stress and Heat Waves [Lennart Olsson (Sweden), Dave Chadee (Trinidad and Tobago), Ove Hoegh-Guldberg (Australia), John Porter (Denmark), Hans-O. Pörtner (Germany), Kirk Smith (USA), Maria Isabel Travasso (Argentina), Petra Tschakert (USA)] Heat waves are periods of abnormally and uncomfortably hot weather during which the risk of heat stress on people and ecosystems is high. The number and intensity of hot days have increased markedly in the last three decades (Coumou et al., 2013) (high confidence). According to WG I, it is likely that the occurrence of heat waves has more than doubled in some locations due to human influence and it is virtually certain that there will be more frequent hot extremes over most land areas in the latter half of the 21st century. Coumou et al. (2013) predicted that, under a medium warming scenario, the number of monthly heat records will be over 12 times more common by the 2040s compared to a non-warming world. In a longer time perspective, if the global mean temperature increases to +10C or more, the habitability of large parts of the tropics and mid-latitudes will be at risk (Sherwood and Huber, 2010). Heat waves affect natural and human systems directly, often with severe losses of lives and assets as a result, and they may act as triggers for tipping points (Hughes et al., 2013). Consequently, heat waves play an important role in several key risks noted in Chapter 19 and CC-KR. Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Economy and Society [Ch 10, 11, 12, 13] Environmental heat stress has already reduced the global labor capacity to 90% in peak months with a further predicted reduction to 80% in peak months by 2050. Under a high warming scenario (RCP8.5), labor capacity is expected to be less than 40% of present day conditions in peak months by 2200 (Dunne et al., 2013). Adaptation costs for securing cooling capacities and emergency shelters during heat waves will be substantial. Heat waves are associated with social predicaments such as increasing violence (Anderson, 2012) as well as overall health and psychological distress and low life satisfaction (Tawatsupa et al., 2012). Impacts are highly differential with disproportional burdens on poor people, elderly people, and those who are marginalized (Wilhelmi et al., 2012). Urban areas are expected to suffer more due to the combined effect of climate and the urban heat island effect (Fischer et al., 2012). In LICs and MICs, adaptation to heat stress is severely restricted for most people in poverty and particularly those who are dependent on working outdoors in agriculture, fisheries, and construction. In small- scale agriculture, women and children are particularly at risk due to the gendered division of labor (Croppenstedt et al., 2013). The expected increase in wildfires as a result of heat waves (Pechony and Shindell, 2010) is a concern for human security, health, and ecosystems. Air pollution from wildfires already causes an estimated 339,000 premature deaths per year worldwide (Johnston et al., 2012). Human Health [Ch 11] Morbidity and mortality due to heat stress is now common all over the world (Barriopedro et al., 2011; Rahmstorf and Coumou, 2011; Nitschke et al., 2011; Diboulo et al., 2012; Hansen et al., 2012). People in physical work are at particular risk as such work produces substantial heat within the body, which cannot be released if the outside temperature and humidity is above certain limits (Kjellstrom et al., 2009). The risk of non-melanoma skin cancer from exposure to UV radiation during summer months increases with temperature (van der Leun, Jan C et al., 2008). Increase in ozone concentrations due to high temperatures affects health (Smith et al., 2010), leading to premature mortality, e.g. cardiopulmonary mortality (Smith et al., 2010). High temperatures are also associated with an increase in air-borne allergens acting as a trigger for respiratory illnesses such as asthma, allergic rhinitis, conjunctivitis, and dermatitis (Beggs, 2010). Ecosystems [Ch 4, 5, 6, 30] Tree mortality is increasing globally (Williams et al., 2012) and can be linked to climate impacts, especially heat and drought (Reichstein et al., 2013), even though attribution to climate change is difficult due to lack of time series and confounding factors. In the Mediterranean region, higher fire risk, longer fire season, and more frequent large, severe fires are expected as a result of increasing heat waves in combination with drought (Duguy et al., 2013), Box 4.2. Marine ecosystem shifts attributed to climate change are often caused by temperature extremes rather than changes in the average (Pörtner and Knust, 2007). During heat exposure near biogeographical limits, even small (<0.5°C) shifts in temperature extremes can have large effects, often exacerbated by concomitant exposures to hypoxia and/or elevated CO2 levels and associated acidification (Hoegh-Guldberg et al., 2007), Figure 6-5, (medium confidence) [Ch 6.3.1, 6.3.5; 30.4; 30.5; CC-MB] Most coral reefs have experienced heat stress sufficient to cause frequent mass coral bleaching events in the last 30 years, sometimes followed by mass mortality (Baker et al., 2008). The interaction of acidification and warming exacerbates coral bleaching and mortality (very high confidence).Temperate seagrass and kelp ecosystems will decline with the increased frequency of heat waves and through the impact of invasive subtropical species (high confidence). [Ch 5, 6, 30.4-30.5, CC-CR, CC-MB] Agriculture [Ch 7] Excessive heat interacts with key physiological processes in crops. Negative yield impacts for all crops past +3C of local warming without adaptation, even with benefits of higher CO2 and rainfall, are expected even in cool environments (Teixeira et al., 2011). For tropical systems where moisture availability or extreme heat limits the length of the growing season, there is a high potential for a decline in the length of the growing season and Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 suitability for crops (medium evidence, medium agreement) (Jones and Thornton, 2009). For example, half of the wheat-growing area of the Indo-Gangetic Plains could become significantly heat-stressed by the 2050s. There is high confidence that high temperatures reduce animal feeding and growth rates (Thornton et al., 2009). Heat stress reduces reproductive rates of livestock (Hansen, 2009), weakens their overall performance (Henry et al., 2012), and may cause mass mortality of animals in feedlots during heat waves (Polley et al., 2013). In the U.S., current economic losses due to heat stress of livestock are estimated at several billion USD annually (St-Pierre et al., 2003). Box CC-HS References Anderson, C.A., 2012: Climate Change and Violence. In: The Encyclopedia of Peace Psychology. [Christie, D.J. (ed.)]. Wiley Online Library Baker, A.C., P.W. Glynn, and B. Riegl, 2008: Climate change and coral reef bleaching: An ecological assessment of long-term impacts, recovery trends and future outlook. Estuarine, Coastal and Shelf Science, 80(4), 435-471. Barriopedro, D., E.M. Fischer, J. Luterbacher, R.M. Trigo, and R. García-Herrera, 2011: The hot summer of 2010: redrawing the temperature record map of Europe. Science, 332(6026), 220-224. Beggs, P.J., 2010: Adaptation to impacts of climate change on aeroallergens and allergic respiratory diseases. International Journal of Environmental Research and Public Health, 7(8), 3006-3021. Coumou, D., A. Robinson, and S. Rahmstorf, 2013: Global increase in record-breaking monthly-mean temperatures. Climatic Change, 118(3-4), 771-782. Croppenstedt, A., M. Goldstein, and N. Rosas, 2013: Gender and agriculture: inefficiencies, segregation, and low productivity traps. The World Bank Research Observer, 28(1), 79-109. Diboulo, E., A. Sie, J. Rocklöv, L. Niamba, M. Ye, C. Bagagnan, and R. Sauerborn, 2012: Weather and mortality: a 10 year retrospective analysis of the Nouna Health and Demographic Surveillance System, Burkina Faso. Global Health Action, 5(19078). Duguy, B., S. Paula, J.G. Pausas, J.A. Alloza, T. Gimeno, and R.V. Vallejo, 2013: Effects of climate and extreme events on wildfire regime and their ecological impacts. In: Regional Assessment of Climate Change in the Mediterranean. Springer, pp. 101-134. Dunne, J.P., R.J. Stouffer, and J.G. John, 2013: Reductions in labour capacity from heat stress under climate warming. Nature Climate Change, published on-line 24 February 2013, 1-4. Fischer, E., K. Oleson, and D. Lawrence, 2012: Contrasting urban and rural heat stress responses to climate change. Geophysical Research Letters, 39(3), L03705. Hansen, J., M. Sato, and R. Ruedy, 2012: Perception of climate change. Proceedings of the National Academy of Sciences, 109(37), E2415- E2423. Hansen, P.J., 2009: Effects of heat stress on mammalian reproduction. Philosophical Transactions of the Royal Society B: Biological Sciences, 364(1534), 3341-3350. Henry, B., R. Eckard, J.B. Gaughan, and R. Hegarty, 2012: Livestock production in a changing climate: adaptation and mitigation research in Australia. Crop and Pasture Science, 63(3), 191-202. Hoegh-Guldberg, O., P. Mumby, A. Hooten, R. Steneck, P. Greenfield, E. Gomez, C. Harvell, P. Sale, A. Edwards, and K. Caldeira, 2007: Coral reefs under rapid climate change and ocean acidification. Science, 318(5857), 1737-1742. Hughes, T.P., S. Carpenter, J. Rockström, M. Scheffer, and B. Walker, 2013: Multiscale regime shifts and planetary boundaries. Trends in Ecology & Evolution, 28(7), 389-395. Johnston, F.H., S.B. Henderson, Y. Chen, J.T. Randerson, M. Marlier, R.S. DeFries, P. Kinney, D.M. Bowman, and M. Brauer, 2012: Estimated global mortality attributable to smoke from landscape fires. Environmental Health Perspectives, 120(5), 695. Jones, P.G. and P.K. Thornton, 2009: Croppers to livestock keepers: livelihood transitions to 2050 in Africa due to climate change. Environmental Science & Policy, 12(4), 427-437. Kjellstrom, T., R. Kovats, S. Lloyd, T. Holt, and R. Tol, 2009: The direct impact of climate change on regional labor productivity. Archives of Environmental & Occupational Health, 64(4), 217-227. Nitschke, M., G.R. Tucker, A.L. Hansen, S. Williams, Y. Zhang, and P. Bi, 2011: Impact of two recent extreme heat episodes on morbidity and mortality in Adelaide, South Australia: a case-series analysis. Environ Health, 10, 42. Pechony, O. and D. Shindell, 2010: Driving forces of global wildfires over the past millennium and the forthcoming century. Proceedings of the National Academy of Sciences, 107(45), 19167-19170. Polley, H.W., D.D. Briske, J.A. Morgan, K. Wolter, D.W. Bailey, and J.R. Brown, 2013: Climate Change and North American Rangelands: Trends, Projections, and Implications. Rangeland Ecology & Management, 66(5), 493-511. Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pörtner, H.O. and R. Knust, 2007: Climate change affects marine fishes through the oxygen limitation of thermal tolerance. Science, 315(5808), 95-97. Rahmstorf, S. and D. Coumou, 2011: Increase of extreme events in a warming world. Proceedings of the National Academy of Sciences, 108(44), 17905-17909. Reichstein, M., M. Bahn, P. Ciais, D. Frank, M.D. Mahecha, S.I. Seneviratne, J. Zscheischler, C. Beer, N. Buchmann, and D.C. Frank, 2013: Climate extremes and the carbon cycle. Nature, 500(7462), 287-295. Sherwood, S.C. and M. Huber, 2010: An adaptability limit to climate change due to heat stress. Proceedings of the National Academy of Sciences, 107(21), 9552-9555. Smith, K.R., M. Jerrett, H.R. Anderson, R.T. Burnett, V. Stone, R. Derwent, R.W. Atkinson, A. Cohen, S.B. Shonkoff, and D. Krewski, 2010: Public health benefits of strategies to reduce greenhouse-gas emissions: health implications of short-lived greenhouse pollutants. The Lancet, 374(9707), 2091-2103. St-Pierre, N., B. Cobanov, and G. Schnitkey, 2003: Economic losses from heat stress by US livestock industries. Journal of Dairy Science, 86, E52-E77. Tawatsupa, B., V. Yiengprugsawan, T. Kjellstrom, and A. Sleigh, 2012: Heat stress, health and well-being: findings from a large national cohort of Thai adults. BMJ Open, 2(6). Teixeira, E.I., G. Fischer, H. van Velthuizen, C. Walter, and F. Ewert, 2011: Global hot-spots of heat stress on agricultural crops due to climate change. Agricultural and Forest Meteorology, 170, 206-215. Thornton, P., J. Van de Steeg, A. Notenbaert, and M. Herrero, 2009: The impacts of climate change on livestock and livestock systems in developing countries: A review of what we know and what we need to know. Agricultural Systems, 101(3), 113-127. van der Leun, Jan C, R.D. Piacentini, and F.R. de Gruijl, 2008: Climate change and human skin cancer. Photochemical & Photobiological Sciences, 7(6), 730-733. Wilhelmi, O., A. de Sherbinin, and M. Hayden, 2012: 12 Exposure to heat stress in urban environments. Ecologies and Politics of Health, 41, 219. Williams, A.P., C.D. Allen, A.K. Macalady, D. Griffin, C.A. Woodhouse, D.M. Meko, T.W. Swetnam, S.A. Rauscher, R. Seager, and H.D. Grissino-Mayer, 2012: Temperature as a potent driver of regional forest drought stress and tree mortality. Nature Climate Change, 3, 292- 297. Box CC-MB. Observed Global Responses of Marine Biogeography, Abundance, and Phenology to Climate Change [Elvira Poloczanska (Australia), Ove Hoegh-Guldberg (Australia), William Cheung (Canada), Hans O. Pörtner (Germany), Michael Burrows (UK)] WGII AR4 presented detection and attribution of a global climate change fingerprint on natural systems (AR4, Ch 1, SPM Figure 1), but studies from marine systems were mostly absent. Since AR4, there has been a rapid increase in studies that focus on climate change impacts on marine species, which represents an opportunity to move from more anecdotal evidence to examining and potentially attributing detected changes within the Ocean to climate change (6.3, Figure MB-1). Recent changes in populations of marine species and the associated shifts in diversity patterns are resulting, at least partly, from climate change-mediated biological responses across ocean regions (6.2, Table 6.7, 30.5) (robust evidence, high agreement, high confidence). Poloczanska et al. (2013) assess a potential pattern in responses of ocean life to recent climate change using a global database of 208 peer-reviewed papers. Observed responses (n=1735) were recorded from 857 species or assemblages across regions and taxonomic groups, from phytoplankton to marine reptiles and mammals (Figure MB-1). Observations were defined as those where the authors of a particular paper assessed the occurrence change in a biological parameter (including distribution, phenology, abundance, demography or community composition) and, if change occurs, the consistency of the change with that expected under climate change. Studies from the peer- reviewed literature were selected using three criteria: (1) authors inferred or directly tested for trends in biological and climatic variables; (2) included data after 1990; and (3) observations spanned at least 19 years, to reduce bias resulting from biological responses to short-term climate variability. [INSERT FIGURE MB-1 HERE Figure MB-1: 1735 observed responses to climate change from 208 single- and multi-species studies. Changes attributed to climate change (blue), inconsistent with climate change (red) and are equivocal (yellow). Each circle Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 represents the centre of a study area. Where points fall on land, it is because they are centroids of distribution that surround an island or peninsula. Pie charts show the proportions within regions bounded by red squares and in the Mediterranean; numbers indicate the total (consistent, opposite or equivocal) observations within each region. Note: 57% of the studies included were published since AR4 (from Poloczanska et al., 2013).] The results of this meta-analysis show that climate change has already had widespread impacts on species distribution, abundance, phenology, and subsequently, species richness and community composition across a broad range of taxonomic groups (plankton to top predators). Of the observations that showed a response in either direction, changes in phenology, distribution and abundance were overwhelmingly (81%) in a direction that was consistent with theoretical responses to climate change (6.2). Knowledge gaps exist, especially in equatorial sub- regions and the Southern Hemisphere (Figure MB-1). The timing of many biological events (phenology) had an earlier onset. For example, over the last 50 years, spring events shifted earlier for many species with an average advancement of 4.4 +/- 0.7 days decade-1 (mean +/- SE) and summer events by 4.4 +/- 1.1 days decade-1 (robust evidence, high agreement, high confidence) (Figure MB-2). Phenological observations included in the study, range from shifts in peak abundance of phytoplankton and zooplankton, to reproduction and migration of invertebrates, fishes and seabirds (6.3.2, 30.5). The distributions of benthic, pelagic and demersal species and communities have shifted by 10s to 1000s of km, although the range shifts have not been uniform across taxonomic groups or ocean regions (6.3.2, 30.5) (robust evidence, high agreement, high confidence). Overall, leading range edges expanded in a poleward direction at 72.0 +/- 13.5 km decade-1 and trailing edges contracted in a poleward direction at 15.8 +/- 8.7 km decade-1 (Figure MB-2) revealing much higher current rates of migration than the potential maximum rates reported for terrestrial species (Figure 4.6) despite slower warming of the Ocean than land surface (WG1 3.2). [INSERT FIGURE MB-2 HERE Figure MB-2. Rates of change in distribution (km decade-1) for marine taxonomic groups, measured at the leading edges (red) and trailing edges (brown). Average distribution shifts calculated using all data, regardless of range location, are in black. Distribution rates have been square-root transformed; standard errors may be asymmetric as a result. Positive distribution changes are consistent with warming (into previously cooler waters, generally poleward). Means +/- standard error are shown, along with number of observations (from Poloczanska et al., 2013).] Poleward distribution shifts have resulted in increased species richness in mid to high latitude regions (Hiddink and ter Hofstede, 2008) and changing community structure (Simpson et al., 2011) (28.2.2). Increases in warm-water components of communities concurrent with regional warming have been observed in mid to high latitude ocean regions including the Bering Sea, Barents Sea, Nordic Sea, North Sea, and Tasman Sea (Box 6.1, 30.5). Observed changes in species composition of catches from 1970 2006 that is partly attributed to long-term ocean warming suggest increasing dominance of warmer water species in sub-tropical and higher latitude regions, and reduction in abundance of sub-tropical species in equatorial waters (Cheung et al., 2013), with implications for fisheries (6.5, 7.4.2, 30.6.2.1) The magnitude and direction of distribution shifts can be related to temperature velocities (i.e., the speed and direction at which isotherms propagate across the Ocean s surface (30.3.1.1, Burrows et al. 2011). Pinsky et al. (2013) showed that shifts in both latitude and depth of benthic fish and crustaceans could be explained by climate velocity with remarkable accuracy, using a database of 128 million individuals across 360 marine taxa from surveys of North American coastal waters conducted over 1968 to 2011. Poloczanska et al. (2013) found that faster distribution shifts generally occur in regions of highest surface temperature velocity, such as the North Sea and sub- Arctic Pacific Ocean. Observed marine species shifts, since approximately 1950s, have generally been able to track observed velocities (Fig MB-3), with phyto- and zooplankton distribution shifts vastly exceeding climate velocities, but with considerable variability within and among taxonomic groups (Poloczanska et al. 2013). Biogeographic shifts are also be influenced by other factors such as nutrient and stratification changes, species interactions, habitat availability and fishing (6.3). Rate and pattern of biogeographic shifts in sedentary organisms and benthic macroalgae are complicated by the influence of local dynamics and topographic features (islands, Subject to Final Copyedit 69 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 channels, coastal lagoons, e.g., of the Mediterranean (Bianchi, 2007), coastal upwelling e.g., Lima et al. (2007)). Geographical barriers constrain range shifts and may cause a loss of endemic species (Ben Rais Lasram et al., 2010), with associated niches filled by alien species, either naturally migrating or artificially introduced (Philippart et al., 2011). Whether marine species can continue to keep pace as warming rates, hence climate velocities, increase (Fig MB-3b) is a key uncertainty. Climate velocities on land are expected to outpace the ability of many terrestrial species to track climate velocities this century (4.3.2.5, Figure 4.6) For marine species, the observed rates of shift are generally much faster than those land for land species, particularly for primary producers and lower trophic levels (Poloczanska et al. 2013). Phyto- and zooplankton communities (excluding larval fish) have extended distributions at remarkable rates (Figure MB-3b), such as in the North-east Atlantic (30.5.1) with implications for marine food webs. Geographical range shifts and depth distribution vary between coexisting marine species (Genner et al., 2004; Perry et al., 2005; Simpson et al., 2011) as a consequence of species-specific thermal window widths and associated vulnerabilities (Figure 6.5). Warming therefore causes differential changes in growth, reproductive success, larval output, early juvenile survival, and recruitment, implying shifts in the relative performance of animal species and, thus, their competitiveness (Pörtner and Farrell, 2008; Figure 6.7A). Such effects may underlie abundance losses or local extinctions, regime shifts between coexisting species, or critical mismatches between predator and prey organisms. Changes in local and regional species richness, abundance, community composition, productivity, energy flows and invasion resistance result. Even among Antarctic stenotherms such differences related to mode of life, phylogeny and associated metabolic capacities exist (6.3.1.4). As a consequence, marine ecosystem functions may be substantially reorganized at the regional scale, potentially triggering a range of cascading effects (Hoegh- Guldberg and Bruno, 2010). A focus on understanding the mechanisms underpinning the nature and magnitude of responses of marine organisms to climate change can help forecast impacts and the associated costs to society and facilitate adaptive management strategies effective in mitigating these impacts (6.3, 6.4). [INSERT FIGURE MB-3 HERE Figure MB-3. A. Rate of climate change for the Ocean (sea surface temperature (SST) °C); B. corresponding climate velocities for the Ocean and median velocity from land (adapted from Burrows et al., 2011); and C. observed rates of displacement of marine taxonomic groups over several decades until 2010. The thin dotted red arrows give an example of interpretation. Rates of climate change of 0.008 °C yr-1 correspond to ca. 2.4 km yr-1median climate velocity in the Ocean. When compared to observed rates of displacement, many marine taxonomic groups have been able to track these velocities, except phyto- and zooplankton where rates of displacement greatly exceed climate velocity. All values are calculated for ocean surface with the exclusion of polar seas (Figure 30-1a). (A) Observed rates of climate change for Ocean SST (Black dotted line) are derived from HadISST1.1 data set, all other rates are calculated based on the average of the CMIP5 climate model ensembles (Table S30-3) for the historical period and for the future based on the four RCP emissions scenarios. Data were smoothed using a 20-year sliding window. (B) Median climate velocity calculated from HadISST1.1 dataset over 1960 2010 using the methods of Burrows et al., 2011. The three axes represent estimated median climate velocities are representative of areas of slow velocities such as Pacific subtropical gyre (STG) system (Purple line), the global Ocean surface (excluding polar seas, Blue line), and areas of high velocities such as the Coral Triangle and North Sea (Orange line). Figure 30-3 shows climate velocities over the ocean surface calculated over 1960 2010. The Red line corresponds to the median rate over global land surface calculated using historical surface temperatures from the CMIP5 model ensemble (Table S30-3). (C) Rates of displacement for marine taxonomic groups estimated by Poloczanska et al. 2013 using published studies (Figure MB-2 Black data set). Note the displacement rates for phytoplankton exceed the axis, so values are given.] Box CC-MB References Ben Rais Lasram, F.B., F. Guilhaumon, C. Albouy, S. Somot, W. Thuiller and D. Mouillot, 2010: The Mediterranean Sea as a cul-de-sac for endemic fishes facing climate change, Global Change Biology, 16, 3233-3245. Bianchi, C.N., 2007: Biodiversity issues for the forthcoming Mediterranean Sea, Hydrobiologia, 580, 7-21. Subject to Final Copyedit 70 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Burrows, M.T., D. S. Schoeman, L.B. Buckley, P.J. Moore, E.S. Poloczanska, K. Brander, K, C.J. Brown, J.F. Bruno, C.M. Duarte, B.S. Halpern, J. Holding, C.V. Kappel, W. Kiessling, M.I. O Connor, J.M. Pandolfi, C. Parmesan, F. Schwing, W.J. Sydeman and A.J. Richardson, 2011; The pace of shifting climate in marine and terrestrial ecosystems, Science, 334,652-655. Cheung, W.W.L., J.L. Sarmiento, J. Dunne, T.L. Frölicher, V. Lam, M.L.D. Palomares, R. Watson and D. Pauly, 2013: Shrinking of fishes exacerbates impacts of global ocean changes on marine ecosystems. Nature Climate Change 3, 254-258. Genner, M.J., D.W. Sims, V.J. Wearmouth, E.J. Southall, A.J. Southward, P.A. Henderson and S.J. Hawkins, 2004: Regional climatic warming drives long-term community changes of British marine fish. Proceedings of the Royal Society of London B: Biological Sciences, 271(1539), 655-661. Hiddink, J. G., and R. ter Hofstede (2008), Climate induced increases in species richness of marine fishes, Global Change Biology, 14, 453-460. Hoegh-Guldberg, O. and J.F. Bruno, 2010: The impact of climate change on the World s marine ecosystems, Science, 328, 1523-1528. Lima, F.P., P.A. Ribeiro, N. Queiroz, S.J. Hawkins and A.M. Santos, 2007; Do distributional shifts of northern and southern species of algae match the warming pattern? Global Change Biology, 13, 2592-2604. Perry, A.L., P.J. Low, J.R. Ellis and J.D. Reynolds, 2005: Climate change and distribution shifts in marine fishes. Science, 308(5730), 1912- 1915. Philippart, C.J.M., R. Anadon, R. Danovaro, J.W. Dippner, K.F. Drinkwater, S.J. Hawkins, T. Oguz, G. O Sullivan and P.C. Reid, 2011: Impacts of climate change on European marine ecosystems: observations, expectations and indicators, Journal of Experimental Marine Biology and Ecology, 400, 52-69. Pinksy, M.L., B. Worm, M.J. Fogarty, J.L. Sarmiento, and S.A. Levin, 2013 Marine taxa track local climate velocities. Science 341, 1239-1242. Pörtner, H.O. and A.P. Farrell, 2008: Ecology: Physiology and climate change. Science, 322(5902), 690-692. Poloczanska, E.S., C.J. Brown, W.J. Sydeman, W. Kiessling, D.S. Schoeman, P.J. Moore, K. Brander, J.F. Bruno, L.B. Buckley, M.T. Burrows, C.M. Duarte, B.S. Halpern, J. Holding, C.V. Kappel, M.I. O Connor, J.M. Pandolfi, C. Parmesan, F. Schwing, S.A.Thompson and A.J. Richardson, 2013: Global imprint of climate change on marine life, Nature Climate Change, published online 4 August 2013, doi: 10.1038/NCLIMATE1958, 7 pp. Simpson, S.D., S. Jennings, M.P. Johnson, J.L. Blanchard, P.J. Schon, D.W. Sims and M.J. Genner, 2011: Continental shelf-wide response of a fish assemblage to rapid warming of the sea, Current Biology, 21, 1565-1570. Box CC-OA. Ocean Acidification [Jean-Pierre Gattuso (France), Peter Brewer (USA), Ove Hoegh-Guldberg (Australia), Joan A. Kleypas (USA), Hans-Otto Pörtner (Germany), Daniela Schmidt (UK)] Anthropogenic ocean acidification and global warming share the same primary cause, which is the increase of atmospheric CO2 (Figure OA-1A; WGI, 2.2.1). Eutrophication, loss of sea ice, upwelling and deposition of atmospheric nitrogen and sulphur all exacerbate ocean acidification locally (5.3.3.6, 6.1.1, 30.3.2.2). [INSERT FIGURE OA-1 HERE Figure OA-1: A: Overview of the chemical, biological, socio-economic impacts of ocean acidification and of policy options (adapted from Turley and Gattuso, 2012). B: Multi-model simulated time series of global mean ocean surface pH (on the total scale) from CMIP5 climate model simulations from 1850 to 2100. Projections are shown for emission scenarios RCP2.6 (blue) and RCP8.5 (red) for the multi-model mean (solid lines) and range across the distribution of individual model simulations (shading). Black (grey shading) is the modelled historical evolution using historical reconstructed forcings. The models that are included are those from CMIP5 that simulate the global carbon cycle while being driven by prescribed atmospheric CO2 concentrations. The number of CMIP5 models to calculate the multi-model mean is indicated for each time period/scenario (WGI AR5 Figure 6.28). C: Effect of near future acidification (seawater pH reduction of 0.5 unit or less) on major response variables estimated using weighted random effects meta-analyses, with the exception of survival which is not weighted (Kroeker et al., 2013). The log- transformed response ratio (LnRR) is the ratio of the mean effect in the acidification treatment to the mean effect in a control group. It indicates which process is most uniformly affected by ocean acidification but large variability exists between species. Significance is determined when the 95% bootstrapped confidence interval does not cross zero. The number of experiments used in the analyses is shown in parentheses. * denotes a statistically significant effect.] Subject to Final Copyedit 71 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chemistry and Projections The fundamental chemistry of ocean acidification is well understood (robust evidence, high agreement). Increasing atmospheric concentrations of CO2 result in an increased flux of CO2 into a mildly alkaline ocean, resulting in a reduction in pH, carbonate ion concentration, and the capacity of seawater to buffer changes in its chemistry (very high confidence). The changing chemistry of the surface layers of the open ocean can be projected at the global scale with high accuracy using projections of atmospheric CO2 levels (Fig. CC-OA-1B). Observations of changing upper ocean CO2 chemistry over time support this linkage (WGI Table 3.2 and Figure 3.18; Figures 30.8, 30.9). Projected changes in open ocean, surface water chemistry for year 2100 based on representative concentration pathways (WGI, Figure 6.28) compared to preindustrial values range from a pH change of -0.14 unit with RCP 2.6 (421 ppm CO2, +1 C, 22% reduction of carbonate ion concentration) to a pH change of -0.43 unit with RCP 8.5 (936 ppm CO2, +3.7 C, 56% reduction of carbonate ion concentration). Projections of regional changes, especially in the highly complex coastal systems (5.3.3.6, 30.3.2.2), in polar regions (WGI 6.4.4), and at depth are more difficult but generally follow similar trends. Biological, Ecological, and Biogeochemical Impacts Investigations of the effect of ocean acidification on marine organisms and ecosystems have a relatively short history, recently analyzed in several metaanalyses (6.3.2.1, 6.3.5.1). A wide range of sensitivities to projected rates of ocean acidification exists within and across diverse groups of organisms, with a trend for greater sensitivity in early life stages (high confidence; 5.4.2.2, 5.4.2.4, 6.3.2). A pattern of positive and negative impacts emerges (high confidence; Fig. OA-1C) but key uncertainties remain in our understanding of the impacts on organisms, life histories and ecosystems. Responses can be influenced, often exacerbated by other drivers, such as warming, hypoxia, nutrient concentration, and light availability (high confidence; 5.4.2.4, 6.3.5). Growth and primary production are stimulated in seagrass and some phytoplankton (high confidence; 5.4.2.3, 6.3.2.2-3, 30.5.6). Harmful algal blooms could become more frequent (limited evidence, medium agreement). Ocean acidification may stimulate nitrogen fixation (limited evidence, low agreement; 6.3.2.2). It decreases the rate of calcification of most, but not all, sea-floor calcifiers (medium agreement, robust evidence) such as reef-building corals (Box CC-CR), coralline algae, bivalves and gastropods reducing the competitiveness with non-calcifiers (5.4.2.2, 5.4.2.4, 6.3.2.5). Ocean warming and acidification promote higher rates of calcium carbonate dissolution resulting in the net dissolution of carbonate sediments and frameworks and loss of associated habitat (medium confidence; 5.4.2.4, 6.3.2.5, 6.3.5.4-5). Some corals and temperate fishes experience disturbances to behavior, navigation and their ability to tell conspecifics from predators (6.3.2.4). However, there is no evidence for these effects to persist on evolutionary timescales in the few groups analyzed (6.3.2). Some phytoplankton and mollusks displayed adaptation to ocean acidification in long-term experiments (limited evidence, medium agreement; 6.3.2.1), indicating that the long-term responses could be less than responses obtained in short-term experiments. However, mass extinctions in Earth history occurred during much slower rates of ocean acidification, combined with other drivers changing, suggesting that evolutionary rates are not fast enough for sensitive animals and plants to adapt to the projected rate of future change (medium confidence; 6.1.2). Projections of ocean acidification effects at ecosystem level are made difficult by the diversity of species-level responses. Differential sensitivities and associated shifts in performance and distribution will change predator-prey relationships and competitive interactions (6.3.2.5, 6.3.5-6), which could impact food webs and higher trophic levels (limited evidence, high agreement). Natural analogues at CO2 vents indicate decreased species diversity, biomass and trophic complexity of communities (Box CC-CR; 5.4.2.3, 6.3.2.5, 30.3.2.2, 30.5). Shifts in community structure have also been documented in regions with rapidly declining pH (5.4.2.2). Due to an incomplete understanding of species-specific responses and trophic interactions the effect of ocean acidification on global biogeochemical cycles is not well understood (limited evidence, low agreement) and represents an important knowledge gap. The additive, synergistic or antagonistic interactions of factors such as temperature, concentrations of oxygen and nutrients, and light are not sufficiently investigated yet. Subject to Final Copyedit 72 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Risks, Socioeconomic Impacts and Costs The risks of ocean acidification to marine organisms, ecosystems, and ultimately to human societies, include both the probability that ocean acidification will affect fundamental physiological and ecological processes of organisms (6.3.2.1), and the magnitude of the resulting impacts on ecosystems and the ecosystem services they provide to society (Box 19-2). For example, ocean acidification under RCP4.5 to RCP8.5 will impact formation and maintenance of coral reefs (high confidence; Box CC-CR, 5.4.2.4) and the goods and services that they provide such as fisheries, tourism and coastal protection (limited evidence, high agreement; Box CC-CR, 6.4.1.1,19.5.2, 27.3.3, 30.5, 30.6). Ocean acidification poses many other potential risks, but these cannot yet be quantitatively assessed due to the small number of studies available, particularly on the magnitude of the ecological and socioeconomic impacts (19.5.2). Global estimates of observed or projected economic costs of ocean acidification do not exist. The largest uncertainty is how the impacts on lower trophic levels will propagate through the food webs and to top predators. However, there are a number of instructive examples that illustrate the magnitude of potential impacts of ocean acidification. A decrease of the production of commercially-exploited shelled mollusks (6.4.1.1) would result in a reduction of US production of 3 to 13% according to the SRES A1FI emission scenario (low confidence). The global cost of production loss of mollusks could be over 100 billion USD by 2100 (limited evidence, medium agreement). Models suggest that ocean acidification will generally reduce fish biomass and catch (low confidence) and that complex additive, antagonistic and/or synergistic interactions will occur with other environmental (warming) and human (fisheries management) factors (6.4.1.1). The annual economic damage of ocean-acidification-induced coral reef loss by 2100 has been estimated, in 2009, to be 870 and 528 billion USD, respectively for the A1 and B2 SRES emission scenarios (low confidence; 6.4.1). Although this number is small compared to global GDP, it can represent a very large GDP loss for the economies of many coastal regions or small islands that rely on the ecological goods and services of coral reefs (25.7.5, 29.3.1.2). Mitigation and Adaptation Successful management of the impacts of ocean acidification includes two approaches: mitigation of the source of the problem (i.e. reduce anthropogenic emissions of CO2), and/or adaptation by reducing the consequences of past and future ocean acidification (6.4.2.1). Mitigation of ocean acidification through reduction of atmospheric CO 2 is the most effective and the least risky method to limit ocean acidification and its impacts (6.4.2.1). Climate geoengineering techniques based on solar radiation management will not abate ocean acidification and could increase it under some circumstances (6.4.2.2). Geoengineering techniques to remove carbon dioxide from the atmosphere could directly address the problem but are very costly and may be limited by the lack of CO2 storage capacity (6.4.2.2). Additionally, some ocean-based approaches, such as iron fertilization, would only re-locate ocean acidification from the upper ocean to the ocean interior, with potential ramifications on deep-water oxygen levels (6.4.2.2; 30.3.2.3 and 30.5.7). A low-regret approach, with relatively limited effectiveness, is to limit the number and the magnitude of drivers other than CO2, such as nutrient pollution (6.4.2.1). Mitigation of ocean acidification at the local level could involve the reduction of anthropogenic inputs of nutrients and organic matter in the coastal ocean (5.3.4.2). Some adaptation strategies include drawing water for aquaculture from local watersheds only when pH is in the right range, selecting for less sensitive species or strains, or relocating industries elsewhere (6.4.2.1). CC-OA References Kroeker K., R.C. Kordas, A. Ryan, I. Hendriks, L.Ramajo, G. Singh, C. Duarte and J.-P. Gattuso, 2013: Impacts of ocean acidification on marine organisms: quantifying sensitivities and interaction with warming. Global Change Biology 19, 1884-1896. Turley C. and J.-P. Gattuso, 2012. Future biological and ecosystem impacts of ocean acidification and their socioeconomic-policy implications. Current Opinion In Environmental Sustainability 4, 278-286. Subject to Final Copyedit 73 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Box CC-PP. Net Primary Production in the Ocean [Philip W. Boyd (New Zealand), Svein Sundby (Norway), Hans-Otto Pörtner (Germany)] Net Primary Production (NPP) is the rate of photosynthetic carbon fixation minus the fraction of fixed carbon used for cellular respiration and maintenance by autotrophic planktonic microbes and benthic plants (6.2.1, 6.3.1). Environmental drivers of NPP include light, nutrients, micronutrients, carbon dioxide, and temperature (Panel A). These drivers in turn, are influenced by oceanic and atmospheric processes, including cloud cover, sea-ice extent, mixing by winds, waves and currents, convection, density stratification, and various forms of upwelling induced by eddies, frontal activity and boundary currents. Temperature has multiple roles as it influences rates of phytoplankton physiology and heterotrophic bacterial recycling of nutrients, in addition to stratification of the water column and sea-ice extent (Panel A). Climate change is projected to strongly impact NPP through a multitude of ways that depend on the regional and local physical settings (WGI, Ch. 3), and on ecosystem structure and functioning (medium confidence, 6.3.4, 6.5.1). The influence of environmental drivers on NPP causes as much as a 10-fold variation in regional productivity: from <50 g C m-2 year-1 in nutrient-poor subtropical waters and light-limited Arctic waters to >> 300 g C m-2 year-1 in productive upwelling regions and highly eutrophic coastal regions (Panel B). The oceans currently provide ~50 x 1015 g C year-1, or about half of global NPP (Field et al. 1998). Global estimates of NPP are mainly obtained from satellite remote-sensing (6.1.2), which provides unprecedented spatial and temporal coverage, and may be validated regionally against oceanic measurements. Observations reveal significant changes in rates of NPP when environmental controls are altered by episodic natural perturbations, such as volcanic eruptions enhancing iron supply, as observed in high-nitrate low-chlorophyll waters of the NE Pacific (Hamme et al., 2010). Climate variability can drive pronounced changes in NPP (Chavez et al., 2011), such as during El Nino to La Nina transitions in Equatorial Pacific, when vertical nutrient and trace element supply are enhanced (Chavez et al., 1999). Multi-year time-series records of NPP have been used to assess spatial trends in NPP in recent decades. Behrenfeld et al. (2006) using satellite data, reported a prolonged and sustained global NPP decrease of 190 x 1012 g C year-1, for the period 1999 to 2005 - an annual reduction of ~0.4 % of global NPP. In contrast, a time-series of directly measured NPP between 1988 to 2007 by Saba et al. (2010) (i.e. in situ incubations using the radiotracer 14C- bicarbonate) revealed an increase (2 % year-1) in NPP for two low latitude open ocean sites. This discrepancy between in situ and remotely-sensed NPP trends points to uncertainties in either the methodology used and/or the extent to which discrete sites are representative of oceanic provinces (Saba et al., 2010, 2011). Modeling studies have subsequently revealed that the <15 year archive of satellite-derived NPP is insufficient to distinguish climate- change mediated shifts in NPP from those driven by natural climate variability (Henson et al., 2010; Beaulieu et al., 2013). Although multidecadal, the available time-series of oceanic NPP measurements are also not of sufficient duration relative to the timescales of climate variability modes (up to 60-70 years for AMO, for example, Figure 6- 1). Recent attempts to synthesize longer (i.e. centennial) records of chlorophyll as a proxy for phytoplankton stocks (e.g., Boyce et al., 2010) have been criticized for relying on questionable linkages between different proxies for chlorophyll over a century of records (e.g., Rykaczewski and Dunne, 2011). Models in which projected climate-change alters the environmental drivers of NPP provide estimates of spatial changes and of the rate of change of NPP. For example, four global coupled climate-ocean biogeochemical Earth System Models (WGI Ch. 6) projected an increase in NPP at high latitudes as a result of alleviation of light and temperature limitation of NPP particularly in Northern and Southern Hemisphere subpolar gyre biomes (Steinacher et al., 2010). However, this regional increase in NPP was more than offset by decreases in NPP at lower latitudes and at mid-latitudes due to the reduced input of macro-nutrients into the photic zone. The reduced mixed-layer depth and reduced rate of circulation may cause a decrease in the flux of macronutrients to the photic zone (Figure 6-2). These changes to oceanic conditions result in a reduction in global mean NPP by 2 to 13% by 2100 relative to 1860 under a high emission scenario (Polovina et al., 2011; SRES A2, between RCP6.0 and RCP8.5). This is consistent with a more recent analysis based on 10 Earth System Models (Bopp et al., 2013), which project decreases in global NPP by 8.6 (+/-7.9), 3.9 (+/-5.7), 3.6 (+/-5.7), 2.0 (+/-4.1) % in the 2090s relative to the 1990s, under the scenarios RCP8.5, RCP6.0, RCP4.5 and RCP2.6, respectively. However, the magnitude of projected changes varies widely Subject to Final Copyedit 74 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 between models (e.g. from 0 to 20% decrease in NPP globally under RCP 8.5). The various models show very large differences in NPP at regional (i.e. provinces, see panel B) scales. Earlier model projections had predicted changes in global NPP from a decrease of > 10% (Field et al., 1998; Boyd and Doney, 2002) to an increase of up to 8.1% under an intermediate scenario (SRES A1B, similar to RCP6.0) (Sarmiento et al., 2004; Schmittner et al., 2008). These projections did not consider the potential contribution of primary production derived from atmospheric nitrogen fixation in tropical and subtropical regions, favoured by increasing stratification and reduced nutrient inputs from mixing. This mechanism is potentially important, although such episodic increases in nitrogen fixation are not sustainable without the presence of excess phosphate (e.g. Moore et al., 2009; Boyd et al., 2010). This may lead to an underestimation of NPP (Mohr et al., 2010; Mulholland et al., 2012; Wilson et al., 2012), however, the extent of such underestimation is unknown (Luo et al., 2012). Care must be taken when comparing global, provincial (e.g. low latitude waters, for example Behrenfeld et al., 2006) and regional trends in NPP derived from observations, as some regions have additional local environmental influences such as enhanced density stratification of the upper ocean from melting sea ice. For example, a longer phytoplankton growing season, due to more sea-ice free days, may have increased NPP (based on a regionally validated time-series of satellite NPP) in Arctic waters (Arrigo and van Dijken, 2011) by an average of 8.1 Tg C year 1 between 1998 and 2009. Other regional trends in NPP are reported in 30.5.1-6. In addition, although future model projections of global NPP from different models (Steinacher et al., 2010; Bopp et al., 2013) are comparable, regional projections from each of the models differ substantially. This raises concerns as to which aspect(s) of the different model NPP parameterizations are responsible for driving regional differences in NPP, and moreover, how accurate model projections are of global NPP. From a global perspective, open ocean NPP will decrease moderately by 2100 under both low (SRES B1 or RCP4.5) and high emission scenarios (A2 or RCP6.0 - 8.5, 6.3.4, 6.5.1, medium confidence), paralleled by an increase in NPP at high latitudes and a decrease in the tropics (medium confidence). However, there is limited evidence and low agreement on the direction, magnitude and differences of a change of NPP in various ocean regions and coastal waters projected by 2100 (low confidence). [INSERT FIGURE PP-1 HERE Figure PP-1: A) Environmental factors controlling Net Primary Production (NPP). NPP is mainly controlled by three basic processes: 1) Light conditions in the surface ocean, i.e. the photic zone where photosynthesis occurs, 2) upward flux of nutrients and micronutrients from underlying waters into the photic zone, 3) Regeneration of nutrients and micronutrients via the breakdown and recycling of organic material before it sinks out of the photic zone. All three processes are influenced by physical, chemical and biological processes and vary across regional ecosystems. In addition, water temperature strongly influences the upper rate of photosynthesis for cells that are resource-replete. Predictions of alteration of primary productivity under climate change depend on correct parameterizations and simulations of each of these variables and processes for each region. B) Annual composite map of global areal NPP rates (derived from MODIS Aqua satellite climatology from 2003-2012; NPP was calculated with the Carbon-based Production Model (CbPM, Westberry et al., 2008)). Overlaid is a grid of (thin black lines) that represent 51 distinct global ocean biogeographical provinces (after Longhurst, 1998 and based on Boyd and Doney, 2002). The characteristics and boundaries of each province are primarily set by the underlying regional ocean physics and chemistry. Figure courtesy of Toby Westberry (OSU) and Ivan Lima (WHOI), satellite data courtesy of NASA Ocean Biology Processing Group.] Box CC-PP References Arrigo, K.R. and G.L. van Dijken, 2011: Secular trends in Arctic Ocean net primary production. Journal of Geophysical Research, 116(C9), C09011. Beaulieu, C., S.A. Henson, J.L. Sarmiento, J.P. Dunne, S.C. Doney, R.R. Rykaczewski and L. Bopp, 2013: Factors challenging our ability to detect long-term trends in ocean chlorophyll. Biogeosciences, 10(4), 2711-2724. Behrenfeld, M.J., R.T. O'Malley, D.A. Siegel, C.R. McClain, J.L. Sarmiento, G.C. Feldman, A.J. Milligan, P.G. Falkowski, R.M. Letelier and E.S. Boss, 2006: Climate-driven trends in contemporary ocean productivity. Nature, 444(7120), 752-755. Subject to Final Copyedit 75 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bopp, L., L. Resplandy, J.D. Orr, D. S.C., J.P. Dunne, M. Gehlen, P. Halloran, C. Heinze, T. Ilyina, R. Séférian, J. Tijiputra and M. Vichi, 2013: Multiple stressors of ocean ecosystems in the 21st century: projections with CMIP5 models. Biogeosciences, 10(10), 6225-6245. Boyce, D.G., M.R. Lewis and B. Worm, 2010: Global phytoplankton decline over the past century. Nature, 466(7306), 591-596. Boyd, P.W. and S.C. Doney, 2002: Modelling regional responses by marine pelagic ecosystems to global climate change. Geophysical Research Letters, 29(16), 1806. Boyd, P.W., R. Strzepek, F.X. Fu and D.A. Hutchins, 2010: Environmental control of open-ocean phytoplankton groups: now and in the future. Limnology and Oceanography, 55(3), 1353-1376. Chavez, F.P., M. Messie and J.T. Pennington, 2011: Marine primary production in relation to climate variability and change. Annual Review of Marine Science, 3(1), 227-260. Chavez, F.P., P.G. Strutton, C.E. Friederich, R.A. Feely, G.C. Feldman, D.C. Foley and M.J. McPhaden, 1999: Biological and chemical response of the equatorial Pacific Ocean to the 1997-98 El Nino. Science, 286(5447), 2126-2131. Field, C.B., M.J. Behrenfeld, J.T. Randerson and P. Falkowski, 1998: Primary production of the biosphere: integrating terrestrial and oceanic components. Science, 281(5374), 237-240. Hamme, R.C., P.W. Webley, W.R. Crawford, F.A. Whitney, M.D. DeGrandpre, S.R. Emerson, C.C. Eriksen, K.E. Giesbrecht, J.F.R. Gower, M.T. Kavanaugh, M.A. Pena, C.L. Sabine, S.D. Batten, L.A. Coogan, D.S. Grundle and D. Lockwood, 2010: Volcanic ash fuels anomalous plankton bloom in subarctic northeast Pacific. Geophysical Research Letters, 37, L19604. Henson, S.A., J.L. Sarmiento, J.P. Dunne, L. Bopp, I. Lima, S.C. Doney, J. John and C. Beaulieu, 2010: Detection of anthropogenic climate change in satellite records of ocean chlorophyll and productivity. Biogeosciences, 7(2), 621-640. Longhurst, A.R., 1998: Ecological Geography of the Sea. Academic Press, San Diego, CA, USA, 560 pp. Luo, Y.-W., S.C. Doney, L.A. Anderson, M. Benavides, I. Berman-Frank, A. Bode, S. Bonnet, K.H. Boström, D. Böttjer, D.G. Capone, E.J. Carpenter, Y.L. Chen, M.J. Church, J.E. Dore, L.I. Falcón, A. Fernández, R.A. Foster, K. Furuya, F. Gómez, K. Gundersen, A.M. Hynes, D.M. Karl, S. Kitajima, R.J. Langlois, J. LaRoche, R.M. Letelier, E. Maranón, D.J. McGillicuddy Jr., P.H. Moisander, C.M. Moore, B. Mourino-Carballido, M.R. Mulholland, J.A. Needoba, K.M. Orcutt, A.J. Poulton, E. Rahav, P. Raimbault, A.P. Rees, L. Riemann, T. Shiozaki, A. Subramaniam, T. Tyrrell, K.A. Turk-Kubo, M. Varela, T.A. Villareal, E.A. Webb, A.E. White, J. Wu and J.P. Zehr, 2012: Database of diazotrophs in global ocean: abundances, biomass and nitrogen fixation rates. Earth System Science Data, 5, 47-106. Mohr, W., T. Grosskopf, D.W.R. Wallace and J. LaRoche, 2010: Methodological underestimation of oceanic nitrogen fixation rates. PLoS ONE, 5(9), e12583. Moore, C.M., M.M. Mills, E.P. Achterberg, R.J. Geider, J. LaRoche, M.I. Lucas, E.L. McDonagh, X. Pan, A.J. Poulton, M.J.A. Rijkenberg, D.J. Suggett, S.J. Ussher and E.M.S. Woodward, 2009: Large-scale distribution of Atlantic nitrogen fixation controlled by iron availability. Nature Geoscience, 2(12), 867-871. Mulholland, M.R., P.W. Bernhardt, J.L. Blanco-Garcia, A. Mannino, K. Hyde, E. Mondragon, K. Turk, P.H. Moisander and J.P. Zehr, 2012: Rates of dinitrogen fixation and the abundance of diazotrophs in North American coastal waters between Cape Hatteras and Georges Bank. Limnology and Oceanography, 57(4), 1067-1083. Polovina, J.J., J.P. Dunne, P.A. Woodworth and E.A. Howell, 2011: Projected expansion of the subtropical biome and contraction of the temperate and equatorial upwelling biomes in the North Pacific under global warming. ICES Journal of Marine Science, 68(6), 986-995. Rykaczewski, R.R. and J.P. Dunne, 2011: A measured look at ocean chlorophyll trends. Nature, 472(7342), E5-E6. Saba, V.S., M.A.M. Friedrichs, D. Antoine, R.A. Armstrong, I. Asanuma, M.J. Behrenfeld, A.M. Ciotti, M. Dowell, N. Hoepffner, K.J.W. Hyde, J. Ishizaka, T. Kameda, J. Marra, F. Mélin, A. Morel, J. O'Reilly, M. Scardi, W.O. Smith Jr., T.J. Smyth, S. Tang, J. Uitz, K. Waters and T.K. Westberry, 2011: An evaluation of ocean color model estimates of marine primary productivity in coastal and pelagic regions across the globe. Biogeosciences, 8(2), 489-503. Saba, V.S., M.A.M. Friedrichs, M.-E. Carr, D. Antoine, R.A. Armstrong, I. Asanuma, O. Aumont, N.R. Bates, M.J. Behrenfeld, V. Bennington, L. Bopp, J. Bruggeman, E.T. Buitenhuis, M.J. Church, A.M. Ciotti, S.C. Doney, M. Dowell, J. Dunne, S. Dutkiewicz, W. Gregg, N. Hoepffner, K.J.W. Hyde, J. Ishizaka, T. Kameda, D.M. Karl, I. Lima, M.W. Lomas, J. Marra, G.A. McKinley, F. Mélin, J.K. Moore, A. Morel, J. O'Reilly, B. Salihoglu, M. Scardi, T.J. Smyth, S.L. Tang, J. Tjiputra, J. Uitz, M. Vichi, K. Waters, T.K. Westberry and A. Yool, 2010: Challenges of modeling depth-integrated marine primary productivity over multiple decades: a case study at BATS and HOT. Global Biogeochemical Cycles, 24, GB3020. Sarmiento, J.L., R. Slater, R. Barber, L. Bopp, S.C. Doney, A.C. Hirst, J. Kleypas, R. Matear, U. Mikolajewicz, P. Monfray, V. Soldatov, S.A. Spall and R. Stouffer, 2004: Response of ocean ecosystems to climate warming. Global Biogeochemical Cycles, 18(3), GB3003. Schmittner, A., A. Oschlies, H.D. Matthews and E.D. Galbraith, 2008: Future changes in climate, ocean circulation, ecosystems, and biogeochemical cycling simulated for a business-as-usual CO2 emission scenario until year 4000 AD. Global Biogeochemical Cycles, 22(1), GB1013. Steinacher, M., F. Joos, T.L. Frölicher, L. Bopp, P. Cadule, V. Cocco, S.C. Doney, M. Gehlen, K. Lindsay, J.K. Moore, B. Schneider and J. Segschneider, 2010: Projected 21st century decrease in marine productivity: a multi-model analysis. Biogeosciences, 7(3), 979-1005. Subject to Final Copyedit 76 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Westberry, T., Behrenfeld, M.J., Siegel, D.A, and Boss, E. 2008. Carbon-based primary productivity modeling with vertically resolved photoacclimation. Global Biogeochemical Cycles, 22(2): GB2024. DOI: 10.1029/2007GB003078 Wilson, S.T., D. Bottjer, M.J. Church and D.M. Karl, 2012: Comparative assessment of nitrogen fixation methodologies, conducted in the oligotrophic North Pacific Ocean. Applied and Environmental Microbiology, 78(18), 6516-6523. Box CC-UP. Uncertain Trends in Major Upwelling Ecosystems [Salvador E. Lluch-Cota (Mexico), Ove Hoegh-Guldberg (Australia), David Karl (USA), Hans O. Pörtner (Germany), Svein Sundby (Norway), Jean-Pierre Gatusso (France)] Upwelling is the vertical transport of cold, dense, nutrient-rich, relatively low-pH and often oxygen-poor waters to the euphotic zone where light is abundant. These waters trigger high levels of primary production and a high biomass of benthic and pelagic organisms. The driving forces of upwelling include wind stress and the interaction of ocean currents with bottom topography. Upwelling intensity also depends on water column stratification. The major upwelling systems of the Planet, the Equatorial Upwelling System (EUS, 30.5.2, Figure 30.1A) and the Eastern Boundary Upwelling Ecosystems (EBUE, 30.5.5, Figure 30.1A), represent only 10% of the ocean surface but contribute nearly 25 % to global fish production (Figure 30.1B, Table S30.1). Marine ecosystems associated with upwelling systems can be influenced by a range of bottom-up trophic mechanisms, with upwelling, transport, and chlorophyll concentrations showing strong seasonal and interannual couplings and variability. These, in turn, influence trophic transfer up the food chain, affecting zooplankton, foraging fish, seabirds and marine mammals. There is considerable speculation as to how upwelling systems might change in a warming and acidifying ocean. Globally, the heat gain of the surface ocean has increased stratification by 4% (WGI 3.2, 3.4.4, 3.8), which means that more wind energy is needed to bring deep waters to the surface. It is as yet unclear to what extent wind stress can offset the increased stratification, due to the uncertainty in wind speed trends (WGI, 3.4.4). In the tropics, observations of reductions in trade winds over several decades contrast more recent evidence indicating their strengthening since the early 1990s (WGI, 9.4.1.3.4). Observations and modelling efforts in fact show diverging trends in coastal upwelling at the eastern boundaries of the Pacific and the Atlantic. Bakun (1990) proposed that the the difference in heat gaining rates between land and ocean causes an increase in the pressure gradient, which results in increased alongshore winds and leads to intensified offshore transport of surface water through Ekman pumping, and the upwelling of nutrient rich, cold waters (Figure CC-UP). Some regional records support this hypothesis, others do not. There is considerable variability in warming and cooling trends over the past decades both within and among systems making it difficult to predict changes in the intensity of all Eastern Boundary Upwelling Ecosystems (30.5.5). Understanding whether upwelling and climate change will impact resident biota in an additive, synergistic or antagonistic manner is important for projections of how ecological goods and services provided for human society will change. Even though upwellings may prove more resilient to climate change than other ocean ecosystems because of their ability to function under extremely variable conditions (Capone and Hutchins, 2013), consequences of their shifts are highly relevant since these are the most biologically active systems in the ocean. Increased upwelling would enhance fisheries yields. However, the export of organic material from surface to deeper layers of the ocean may increase and stimulate its decomposition by microbial activity, thereby enhancing oxygen depletion and CO2 enrichment in deeper water layers. Once this water returns to the surface through upwelling benthic and pelagic coastal communities will be exposed to acidified and deoxygenated water which may combine with anthropogenic impact to negatively affect marine biota and ecosystem structure of the upper ocean (high confidence, 6.3.2, 6.3.3; 30.3.2.2, 30.3.2.3). Extreme hypoxia may result in abnormal mortalities of fishes and invertebrates (Keller et al., 2010), reduce the fisheries catch potential and impact aquaculture in coastal areas (5.4.3.3, 6.3.7, 30.5.1.1.2, 30.5.5.1.3, Barton et al., 2012). Shifts in upwelling also coincide with an apparent increase in the frequency of submarine eruptions of methane and hydrogen sulphide gas, caused by enhanced formation and sinking of phytoplankton biomass to the hypoxic or anoxic sea floor . This combination of factors has been implicated in the extensive mortality of coastal fishes and invertebrates (Bakun and Weeks, 2004), resulting in significant reductions Subject to Final Copyedit 77 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 in fishing productivity, such as Cape hake (Merluccius capensis), Namibia s most valuable fishery (Hamukuaya et al., 1998). Reduced upwelling would also reduce the productivity of important pelagic fisheries, such as for sardines, anchovies and mackerel, with major consequences for the economies of several countries (6.4.1, Chp 7, Figure 30.1A, B, Table S30.1). However, under projected scenarios of reduced upward supply of nutrients due to stratification of the open ocean , upwelling of both nutrients and trace elements may become increasingly important to maintaining upper ocean nutrient and trace metal inventories. It has been suggested that upwelling areas may also increase nutrient content and productivity under enhanced stratification, and that upwelled and partially denitrified waters containing excess phosphate may select for N2-fixing microorganisms (Deutsch et al., 2007; Deutsch and Weber, 2012), but field observations of N2 fixation in these regions have not supported these predictions (Fernandez et al., 2011; Franz et al., 2012). The role of this process in global primary production thus needs to be validated (low confidence). The central question therefore is whether or not upwelling will intensify, and if so, whether the effects of intensified upwelling on O2 and CO2 inventories will outweigh its benefits for primary production and associated fisheries and aquaculture (low confidence). In any case increasing atmospheric CO2 concentrations will equilibrate with upwelling waters that may cause them to become more corrosive, depending upon pCO2 of the upwelled water, and potentially increasingly impact the biota of Eastern Boundary Upwelling Ecosystems. [INSERT FIGURE UP-1 HERE Figure UP-1: Upper panel: Schematic hypothetic mechanism of increasing coastal wind-driven upwelling at eastern boundary systems, where differential warming rates between land and ocean results in increased land-ocean pressure gradients (1) that produce stronger alongshore winds (2) and offshore movement of surface water through Ekman transport (3), and increased upwelling of deep cold nutrient rich waters to replace it (4). Lower panel: potential consequences of climate change in upwelling systems. Increasing stratification and uncertainty in wind stress trends result in uncertain trends in upwelling. Increasing upwelling may result in higher input of nutrients to the euphotic zone, and increased primary production, which in turn may enhance pelagic fisheries, but also decreased coastal fisheries due to an augmented exposure of coastal fauna to hypoxic, low pH waters. Decreased upwelling may result in lower primary production in these systems with direct impacts on pelagic fisheries productivity.] Box CC-UP References Bakun, A., 1990: Global climate change and intensification of coastal ocean upwelling, Science, 247(4939), 198-201. Bakun, A. and S.J. Weeks, 2004: Greenhouse gas buildup, sardines, submarine eruptions and the possibility of abrupt degradation of intense marine upwelling ecosystems. Ecology Letters, 7(11), 1015-1023. Barton, A., B. Hales, G.G. Waldbusser, C. Langdon, R.A. Feely, 2012: The Pacific oyster, Crassostrea gigas, shows negative correlation to naturally elevated carbon dioxide levels: Implications for near-term ocean acidification effects, Limnology and Oceanography, 57(3): 698- 710. Capone, D.G. and D.A. Hutchins, 2013: Microbial biogeochemistry of coastal upwelling regimes in a changing ocean. Nature geoscience, 711- 717. Deutsch, C. and T. Weber, 2012: Nutrient ratios as a tracer and driver of ocean biogeochemistry. Annual Review of Marine Science, 4, 113-114. Deutsch, C., J.L. Sarmiento, D.M. Sigman, N. Gruber and J.P. Dunne, 2007: Spatial coupling of nitrogen inputs and losses in the ocean. Nature, 445(7124), 163-167. Fernandez, C., L. Farías and O. Ulloa, 2011: Nitrogen fixation in denitrified marine waters. PLoS ONE, 6(6), e20539. Franz, J., G. Krahmann, G. Lavik, P. Grasse, T. Dittmar and U. Riebesell, 2012: Dynamics and stoichiometry of nutrients and phytoplankton in waters influenced by the oxygen minimum zone in the eastern tropical Pacific. Deep Sea Research Part I: Oceanographic Research Papers, 62, 20-31. Hamukuaya, H., M.J. O'Toole and P.M.J. Woodhead, 1998: Observations of severe hypoxia and offshore displacement of Cape hake over the Namibian shelf in 1994. South African Journal of Marine Science, 19(1), 57-59. Keller, AA, Simon V, Chan F, Wakefield WW, Clarke ME, et al., 2010: Demersal fish and invertebrate biomass in relation to an offshore hypoxic zone along the US West Coast. Fisheries Oceanography 19:76 87. Subject to Final Copyedit 78 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 References Adams, C. M., E. Hernandez, and J. C. Cato (2004), The economic significance of the Gulf of Mexico related to population, income, employment, minerals, fisheries and shipping, Ocean & Coastal Management, 47(11-12), 565-580. Adjeroud, M., F. Michonneau, P. Edmunds, Y. Chancerelle, T. de Loma, L. Penin, L. Thibaut, J. Vidal-Dupiol, B. Salvat, and R. Galzin (2009), Recurrent disturbances, recovery trajectories, and resilience of coral assemblages on a South Central Pacific reef, Coral Reefs, 28(3), 775-780. Al-Rashidi, T. B., H. I. El-Gamily, C. L. Amos, and K. A. Rakha (2009), Sea surface temperature trends in Kuwait Bay, Arabian Gulf, Natural Hazards, 50(1), 73-82. Albright, R., and C. Langdon (2011), Ocean acidification impacts multiple early life history processes of the Caribbean coral Porites astreoides, Global Change Biol, 17(7), 2478-2487. Albright, R., B. Mason, M. Miller, and C. Langdon (2010), Ocean acidification compromises recruitment success of the threatened Caribbean coral Acropora palmata, Proceedings of the National Academy of Sciences of the United States of America, 107(47), 20400-20404. Aleksandrov, S. V., N. N. Zhigalova, and A. S. Zezera (2009), Long-term dynamics of zooplankton in the southeastern Baltic Sea, Russ J Mar Biol +, 35(4), 296-304. Alheit, J. (2009), Consequences of regime shifts for marine food webs, International Journal of Earth Sciences, 98(2), 261-268. Alheit, J., and A. Bakun (2010), Population synchronies within and between ocean basins: Apparent teleconnections and implications as to physical-biological linkage mechanisms, J Marine Syst, 79(3-4), 267-285. Allen, R., and A. Bergin (2009), Ocean acidification: An emerging Australian environmental security challenge, Australian Journal of Maritime and Ocean Affairs, 1(2), 49. Allison, E. H., A. L. Perry, M. C. Badjeck, W. N. Adger, K. Brown, D. Conway, A. S. Halls, G. M. Pilling, J. D. Reynolds, and N. L. Andrew (2009), Vulnerability of national economies to the impacts of climate change on fisheries, Fish Fish, 10(2), 173-196 Alonso-Pérez, S., E. Cuevas, and X. Querol (2011), Objective identification of synoptic meteorological patterns favouring African dust intrusions into the marine boundary layer of the subtropical eastern north Atlantic region, Meteorology and Atmospheric Physics, 113(3-4), 109-124. Amelung, B., S. Nicholls, and D. Viner (2007), Implications of global climate change for tourism flows and seasonality, Journal of Travel Research, 45(3), 285-296. Andersson, A. J., and D. Gledhill (2013), Ocean acidification and coral reefs: Effects on breakdown, dissolution, and net ecosystem calcification, Annual Review of Marine Science, 5, 321-348. Andrews, O. D., N. L. Bindoff, P. R. Halloran, T. Ilyina, and C. Le Quéré (2013), Detecting an external influence on recent changes in oceanic oxygen using an optimal fingerprinting method, Biogeosciences, 10(3), 1799-1813. Antonov, J. I., D. Seidov, T. P. Boyer, R. A. Locarnini, A. V. Mishonov, H. E. Garcia, O. K. Baranova, M. M. Zweng, and D. R. Johnson (2010), World Ocean Atlas 2009, Volume 2: Salinity. S. Levitus, Ed, in NOAA Atlas NESDIS, edited, p. 184, U.S. Government Printing Office, Washington DC. Archer, D., B. Buffett, and V. Brovkin (2009), Ocean methane hydrates as a slow tipping point in the global carbon cycle, Proceedings of the National Academy of Sciences of the United States of America, 106(49), 20596- 20601. Arístegui, J., E. D. Barton, X. A. Álvarez-Salgado, A. M. P. Santos, F. G. Figueiras, S. Kifani, S. Hernández-León, E. Mason, E. Machú, and H. Demarcq (2009), Sub-regional ecosystem variability in the Canary Current upwelling, Progress in Oceanography, 83(1), 33-48. Arrigo, K. R., G. van Dijken, and S. Pabi (2008), Impact of a shrinking Arctic ice cover on marine primary production, Geophysical Research Letters, 35(19), L19603. Arthun, M., T. Eldevik, L. H. Smedsrud, O. Skagseth, and R. B. Ingvaldsen (2012), Quantifying the influence of Atlantic heat on Barents Sea ice variability and retreat, Journal of Climate, 25(13), 4736-4743. Ashok, K., S. K. Behera, S. A. Rao, H. Y. Weng, and T. Yamagata (2007), El Nino Modoki and its possible teleconnection, J Geophys Res-Oceans, 112(C11), C11007. Astthorsson, O. S., H. Valdimarsson, A. Gudmundsdottir, and G. J. Óskarsson (2012), Climate-related variations in the occurrence and distribution of mackerel (Scomber scombrus) in Icelandic waters, ICES Journal of Marine Science, 69(7), 1289-1297. Subject to Final Copyedit 79 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ateweberhan, M., and T. R. McClanahan (2010), Relationship between historical sea-surface temperature variability and climate change-induced coral mortality in the western Indian Ocean, Mar Pollut Bull, 60(7), 964-970. Ateweberhan, M., T. R. McClanahan, N. A. J. Graham, and C. R. C. Sheppard (2011), Episodic heterogeneous decline and recovery of coral cover in the Indian Ocean, Coral Reefs, 30(3), 739-752. Auad, G., D. Roemmich, and J. Gilson (2011), The California Current System in relation to the Northeast Pacific Ocean circulation, Progress in Oceanography, 91, 576-592. Ayón, P., S. Purca, and R. Guevara-Carrasco (2004), Zooplankton volume trends off Peru between 1964 and 2001, ICES Journal of Marine Science, 61(4), 478-484. Baker-Austin, C., J. A. Trinanes, N. G. H. Taylor, R. Hartnell, A. Siitonen, and J. Martinez-Urtaza (2012), Emerging Vibrio risk at high latitudes in response to ocean warming, Nature Climate Change, 3(1), 73-77. Baker, A., P. W. Glynn, and B. Riegl (2008), Climate change and coral reef bleaching: An ecological assessment of long-term impacts, recovery trends and future outlook, Estuar Coast Shelf S, 80(4), 435-471. Bakun, A. (1990), Global climate change and intensification of coastal ocean upwelling, Science, 247(4939), 198- 201. Bakun, A., D. B. Field, A. Redondo-Rodriguez, and S. J. Weeks (2010), Greenhouse gas, upwelling-favorable winds, and the future of coastal ocean upwelling ecosystems, Global Change Biol, 16(4), 1213-1228. Bank, W., and FAO (2008), The Sunken Billions: The Economic Justification for Fisheries Reform, edited, p. 130, World Bank, Washington, DC. Banks, S. A. M. C., S. D. Ling, C. R. Johnson, M. P. Piggott, J. E. Williamson, and L. B. Beheregaray (2010), Genetic structure of a recent climate change driven range extension, Molecular Ecology, 19(10), 2011-2024. Banón, R. D. (2009), Variacións na diversidade e abundancia ictiolóxica marina en Galicia por efectos do cambio climático, Evidencias e impacto do Cambio Climático en Galicia. X. de Galicia. Santiago Composten, Xunta de Galicia, 355-372. Barber, R. T. (1988), Ocean basin ecosystems, in Concepts of Ecosystem Ecology, edited by L. R. Pomeroy and J. J. Alberts, pp. 171-193, Springer-Verlag, New York. Barber, R. T., J. W. Murray, and J. J. McCarthy (1994), Biogeochemical interactions in the equatorial Pacific, Ambio, 23(1), 62-66. Barlow, J., M. Kahru, and B. Mitchell (2008), Cetacean biomass, prey consumption, and primary production requirements in the California Current ecosystem, Marine Ecology Progress Series, 371, 285-295. Barton, A., B. Hales, G. G. Waldbusser, C. Langdon, and R. A. Feely (2012), The Pacific oyster, Crassostrea gigas, shows negative correlation to naturally elevated carbon dioxide levels: Implications for near-term ocean acidification effects, Limnol Oceanogr, 57(3), 698-710. Barton, E. D., D. B. Field, and C. Roy (2013), Canary current upwelling: More or less?, Progress in Oceanography, 116, 167-178. Batten, S. D., and A. W. Walne (2011), Variability in northwards extension of warm water copepods in the NE Pacific, J Plankton Res, 33(11), 1643-1653. Bauman, A. G., J. A. Burt, D. A. Feary, E. Marquis, and P. Usseglio (2010), Tropical harmful algal blooms: An emerging threat to coral reef communities?, Mar Pollut Bull, 60(11), 2117-2122. Beare, D., F. Burns, E. Jones, K. Peach, and D. Reid (2005), Red mullet migration into the northern North Sea during late winter, Journal of Sea Research, 53(3), 205-212. Beare, D. J., F. Burns, A. Greig, E. G. Jones, K. Peach, M. Kienzle, E. McKenzie, and D. G. Reid (2004), Long-term increases in prevalence of North Sea fishes having southern biogeographic affinities, Marine Ecology Progress Series, 284, 269-278. Beaudoin, Y., and L. Pendleton (2012), Why value the oceans?, A discussion paper prepared by UNEP/GRID- Arendal and Duke University's Nicholas Institute for Environmental Policy Solutions in collaboration with the UNEP TEEB office and the UNEP Regional Seas Program, 33. Beaugrand, G. (2004), The North Sea regime shift: evidence, causes, mechanisms and consequences, Progress in Oceanography, 60(2), 245-262. Beaugrand, G. (2009), Decadal changes in climate and ecosystems in the North Atlantic Ocean and adjacent seas, Deep-Sea Res Pt Ii, 56(8-10), 656-673. Beaugrand, G., and P. C. Reid (2003), Long-term changes in phytoplankton, zooplankton and salmon related to climate, Global Change Biol, 9(6), 801-817. Beaugrand, G., and R. R. Kirby (2010), Spatial changes in the sensitivity of Atlantic cod to climate-driven effects in the plankton, Clim Res, 41(1), 15-19. Subject to Final Copyedit 80 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Beaugrand, G., M. Edwards, and L. Legendre (2010), Marine biodiversity, ecosystem functioning, and carbon cycles, Proceedings of the National Academy of Sciences of the United States of America, 107(22), 10120- 10124. Beaugrand, G., P. C. Reid, F. Ibanez, J. A. Lindley, and M. Edwards (2002), Reorganization of North Atlantic marine copepod biodiversity and climate, Science, 296(5573), 1692-1694. Beaugrand, G., K. M. Brander, J. A. Lindley, S. Souissi, and P. C. Reid (2003), Plankton effect on cod recruitment in the North Sea, Nature, 426(6967), 661-664. Bednar¹ek, N., G. A. Tarling, D. C. E. Bakker, S. Fielding, E. M. Jones, H. J. Venables, P. Ward, A. Kuzirian, B. Lézé, and R. A. Feely (2012), Extensive dissolution of live pteropods in the Southern Ocean, Nature Geoscience, 5, 881-885. Belkin, I. M. (2009), Rapid warming of large marine ecosystems, Progress in Oceanography, 81(1-4), 207-213. Bell, J. D., J. E. Johnson, and A. J. Hobday (2011a), Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, 925 pp., Secretariat of the Pacific Community, Noumea, New Caledonia. Bell, J. D., C. Reid, M. J. Batty, P. Lehodey, L. Rodwell, A. J. Hobday, J. E. Johnson, and A. Demmke (2013a), Effects of climate change on oceanic fisheries in the tropical Pacific: implications for economic development and food security, Climatic Change, 119, 1-14. Bell, J. D., et al. (2011b), Implications of climate change for contributions by fisheries and aquaculture to Pacific Island economies and communities, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. J. Hobday, pp. 733-802, Secretariat of the Pacific Community, Noumea, New Caledonia. Bell, J. D., et al. (2013b), Mixed responses of tropical Pacific fisheries and aquaculture to climate change, Nature Climate Change, 3, 591-599. Ben Rais Lasram, F., and D. Mouillot (2008), Increasing southern invasion enhances congruence between endemic and exotic Mediterranean fish fauna, Biological Invasions, 11(3), 697-711. Bensoussan, N., J. C. Romano, J. G. Harmelin, and J. Garrabou (2010), High resolution characterization of northwest Mediterranean coastal waters thermal regimes: To better understand responses of benthic communities to climate change, Estuar Coast Shelf S, 87(3), 431-441. Berke, S. K., A. R. Mahon, F. P. Lima, K. M. Halanych, D. S. Wethey, and S. A. Woodin (2010), Range shifts and species diversity in marine ecosystem engineers: patterns and predictions for European sedimentary habitats, Global Ecology and Biogeography, 19(2), 223-232. Berumen, M. L., and M. S. Pratchett (2006), Recovery without resilience: persistent disturbance and long-term shifts in the structure of fish and coral communities at Tiahura Reef, Moorea, Coral Reefs, 25(4), 647-653. Bethoux, J., B. Gentili, J. Raunet, and D. Tailliez (1990), Warming trend in the western Mediterranean deep water, Nature, 347, 660-662. Beukema, J. J., R. Dekker, and J. M. Jansen (2009), Some like it cold: populations of the tellinid bivalve Macoma balthica (L.) suffer in various ways from a warming climate, Marine Ecology Progress Series, 384, 135-145. Bi, H. S., W. T. Peterson, and P. T. Strub (2011), Transport and coastal zooplankton communities in the northern California Current system, Geophysical Research Letters, 38, L12607. Bianchi, C. N. (2007), Biodiversity issues for the forthcoming tropical Mediterranean Sea, Hydrobiologia, 580, 7- 21. Blanchard, J. L., S. Jennings, R. Holmes, J. Harle, G. Merino, J. I. Allen, J. Holt, N. K. Dulvy, and M. Barange (2012), Potential consequences of climate change for primary production and fish production in large marine ecosystems, Philosophical Transactions of the Royal Society B: Biological Sciences, 367(1605), 2979-2989. Block, B. A., et al. (2011), Tracking apex marine predator movements in a dynamic ocean, Nature, 475, 86-90. Bode, A., R. Anadón, X. A. G. Morán, E. Nogueira, E. Teira, and M. Varela (2011), Decadal variability in chlorophyll and primary production off NW Spain, Clim Res, 48(2), 293-305. Bograd, S. J., C. G. Castro, E. Di Lorenzo, D. M. Palacios, H. Bailey, W. Gilly, and F. P. Chavez (2008), Oxygen declines and the shoaling of the hypoxic boundary in the California Current, Geophysical Research Letters, 35(12), L12607. Bonfils, C., and B. D. Santer (2011), Investigating the possibility of a human component in various pacific decadal oscillation indices, Clim Dynam, 37(7), 1457-1468. Bonnet, D., et al. (2005), An overview of Calanus helgolandicus ecology in European waters, Progress in Oceanography, 65(1), 1-53. Subject to Final Copyedit 81 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Borgerson, S. G. (2008), Arctic meltdown - The economic and security implications of global warming, Foreign Aff, 87(2), 63-77. Bostock, J., et al. (2010), Aquaculture: global status and trends, Philosophical Transactions of the Royal Society B: Biological Sciences, 365, 2897-2912. Boyer, T. P. (2005), Linear trends in salinity for the World Ocean, 1955 1998, Geophysical Research Letters, 32(1), L01604. Boyle, A. (2012), Law of the Sea perspectives on climate change, The International Journal of Marine and Coastal Law, 27(4), 831-838. Bozkurt, D., and O. L. Sen (2011), Precipitation in the Anatolian Peninsula: sensitivity to increased SSTs in the surrounding seas, Clim Dynam, 36(3-4), 711-726. Brevik, E. C., and J. A. Homburg (2004), A 5000 year record of carbon sequestration from a coastal lagoon and wetland complex, Southern California, USA, Catena, 57(3), 221-232. Brown, B. E., R. P. Dunne, N. Phongsuwan, and P. J. Somerfield (2011), Increased sea level promotes coral cover on shallow reef flats in the Andaman Sea, eastern Indian Ocean, Coral Reefs, 30(4), 867-878. Brown, C. W., and J. A. Yoder (1994), Coccolithophorid blooms in the global ocean, J Geophys Res-Oceans, 99(C4), 7467-7482. Brown, J. H., J. F. Gillooly, A. P. Allen, V. M. Savage, and G. B. West (2004), Toward a metabolic theory of ecology, Ecology, 85(7), 1771-1789. Brundrit, G. B. (1995), Trends in Southern african sea level: Statistical analysis and interpretation, South African Journal of Marine Science, 16, 9-17. Bruno, J. F., and E. R. Selig (2007), Regional decline of coral cover in the Indo-Pacific: timing, extent, and subregional comparisons, Plos One, 2(8), e711. Buesseler, K. O., et al. (2007), Revisiting carbon flux through the ocean's twilight zone, Science, 316(5824), 567- 570. Buongiorno Nardelli, B., S. Colella, R. Santoleri, M. Guarracino, and A. Kholod (2010), A re-analysis of Black Sea surface temperature, J Marine Syst, 79(1-2), 50-64. Burke, L. M., and J. Maidens (2004), Reefs at Risk in the Caribbean, 84 pp., World Resources Institute, Washington DC. Burke, L. M., L. Selig, and M. Spalding (2002), Reefs at Risk in Southeast Asia, 76 pp., World Resources Institute, Washington DC. Burke, L. M., K. Reytar, M. Spalding, and A. Perry (2011), Reefs at Risk Revisited, 130 pp., World Resources Institute, Washington DC. Burrows, M. T., et al. (2011), The pace of shifting climate in marine and terrestrial ecosystems, Science, 334(6056), 652-655. Busby, J. W. (2009), The climate-security connection: What it means for the poor, in Climate Change and Global Poverty: A Billion Lives in the Balance, edited by L. Brainard, A. Jones and N. Purvis, pp. 155-180, Brookings Institute Press, Washington DC. Byrne, R. H., S. Mecking, R. A. Feely, and X. Liu (2010), Direct observations of basin-­ wide acidification of the North Pacific Ocean, Geophysical Research Letters, 37(2), L02601. Cai, R., and H. Tan (2010), Influence of interdecadal climate variation over East Asia on offshore ecological system of China, Journal of Oceanography in Taiwan Strait, 29(2), 173-183 (in Chinese, with English abstract). Cai, R., J. Chen, and R. Huang (2006), The response of marine environment in the offshore area of China and its adjacent ocean to recent global climate change, Chinese Journal of Atmospheric Sciences, 30(5), 1019-1033 (in Chinese, with English abstract). Cai, R., Q. Zhang, and Q. Qi (2009), Spatial and temporal oscillation and long-term variation in sea surface temperature field of the South China Sea, Journal of Oceanography in Taiwan Strait, 29(2), 173-183 (in Chinese, with English abstract). Cai, R., J. Chen, and H. Tan (2011), Variations of the sea surface temperature in the offshore area of China and their relationship with the East Asian monsoon under the global warming, Climatic and Environmental Research, 16(1), 94-104 (in Chinese, with English abstract). Cai, W.-Y., et al. (2012), More extreme swings of the South Pacific convergence zone due to greenhouse warming, Nature 488, 365-369. Cai, W. (2006), Antarctic ozone depletion causes an intensification of the Southern Ocean super-gyre circulation, Geophysical Research Letters, 33(3), L03712. Subject to Final Copyedit 82 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Cai, W. J., and T. Cowan (2007), Trends in Southern Hemisphere circulation in IPCC AR4 models over 1950-99: Ozone depletion versus greenhouse forcing, Journal of Climate, 20(4), 681-693. Caldeira, K., and M. E. Wickett (2003), Oceanography: Anthropogenic carbon and ocean pH, Nature, 425(6956), 365-365. Callaghan, J., and S. B. Power (2011), Variability and decline in the number of severe tropical cyclones making land-fall over eastern Australia since the late nineteenth century, Clim Dynam, 37(3), 647-662. Calvo, E., R. Simó, R. Coma, M. Ribes, J. Pascual, A. Sabatés, J. M. Gili, and C. Pelejero (2011), Effects of climate change on Mediterranean marine ecosystems: the case of the Catalan Sea, Clim Res, 50(1), 1-29. Cantin, N. E., A. L. Cohen, K. B. Karnauskas, A. M. Tarrant, and D. C. McCorkle (2010), Ocean warming slows coral growth in the central Red Sea, Science, 329(5989), 322-325. Carrigan, A. D., and M. L. Puotinen (2011), Assessing the potential for tropical cyclone induced sea surface cooling to reduce thermal stress on the world's coral reefs, Geophysical Research Letters, 38(23), L23604. Casini, M., G. Kornilovs, M. Cardinale, C. Möllmann, W. Grygiel, P. Jonsson, T. Raid, J. Flinkman, and V. Feldman (2011), Spatial and temporal density dependence regulates the condition of the central Baltic Sea clupeids: compelling evidence using an extensive international acoustic survey, Population Ecology, 53(4), 511-523. Cerrano, C., G. Bavestrello, C. Bianchi, R. Cattaneo-­ Vietti, S. Bava, C. Morganti, C. Morri, P. Picco, G. Sara, and S. Schiaparelli (2000), A catastrophic mass-mortality episode of gorgonians and other organisms in the Ligurian Sea (North-western Mediterranean), summer 1999, Ecology Letters, 3(4), 284-293. Chan, F., J. A. Barth, J. Lubchenco, A. Kirincich, H. Weeks, W. T. Peterson, and B. A. Menge (2008), Emergence of anoxia in the California current large marine ecosystem, Science, 319(5865), 920. Chan, W. H., Y. L. Mak, J. J. Wu, L. Jin, W. H. Sit, J. C. Lam, Y. Sadovy de Mitcheson, L. L. Chan, P. K. Lam, and M. B. Murphy (2011), Spatial distribution of ciguateric fish in the Republic of Kiribati, Chemosphere, 84(1), 117-123. Chang, Y., M. A. Lee, K. T. Lee, and K. T. Shao (2013), Adaptation of fisheries and mariculture management to extreme oceanic environmental changes and climate variability in Taiwan, Mar Policy, 38, 476-482. Chassot, E., F. Mélin, O. Le Pape, and D. Gascuel (2007), Bottom-up control regulates fisheries production at the scale of eco-regions in European seas, Marine Ecology Progress Series, 343, 45-55. Chassot, E., S. Bonhommeau, N. K. Dulvy, F. Mélin, R. Watson, D. Gascuel, and O. Le Pape (2010), Global marine primary production constrains fisheries catches, Ecology Letters, 13(4), 495-505. Chavez, F. P., and M. Messie (2009), A comparison of Eastern Boundary Upwelling Ecosystems, Progress in Oceanography, 83(1-4), 80-96. Chavez, F. P., M. Messie, and J. T. Pennington (2011), Marine primary production in relation to climate variability and change, Annual Review of Marine Science, 3, 227-260. Chavez, F. P., J. Ryan, S. E. Lluch-Cota, and C. M. Niquen (2003), From anchovies to sardines and back: multidecadal change in the Pacific Ocean, Science, 299(5604), 217-221. Checkley, D. M., and J. A. Barth (2009), Patterns and processes in the California Current System, Progress in Oceanography, 83(1-4), 49-64. Chelton, D. B., P. A. Bernal, and J. A. McGowan (1982), Large-scale interannual physical and biological interaction in the California Current, Journal of Marine Research, 40(4), 1095-1125. Chen, T. R., K. F. Yu, Q. Shi, S. Li, G. J. Price, R. Wang, M. X. Zhao, T. G. Chen, and J. X. Zhao (2009), Twenty- five years of change in scleractinian coral communities of Daya Bay (northern South China Sea) and its response to the 2008 AD extreme cold climate event, Chinese Science Bulletin, 54(12), 2107-2117. Cheng, X. H., and Y. Q. Qi (2007), Trends of sea level variations in the South China Sea from merged altimetry data, Global and Planetary Change, 57(3-4), 371-382. Cheung, W. W. L., J. Dunne, J. L. Sarmiento, and D. Pauly (2011), Integrating ecophysiology and plankton dynamics into projected maximum fisheries catch potential under climate change in the Northeast Atlantic, ICES Journal of Marine Science, 68(6), 1008-1018. Cheung, W. W. L., J. K. Pinnegar, G. Merino, M. C. Jones, and M. Barange (2012), Review of climate change impacts on marine fisheries in the UK and Ireland, Aquat. Conserv.-Mar. Freshw. Ecosyst., 22, 368-388. Cheung, W. W. L., V. W. Y. Lam, J. L. Sarmiento, K. Kearney, R. Watson, D. Zeller, and D. Pauly (2010), Large- scale redistribution of maximum fisheries catch potential in the global ocean under climate change, Global Change Biol, 16(1), 24-35. Subject to Final Copyedit 83 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chhak, K. C., E. Di Lorenzo, N. Schneider, and P. F. Cummins (2009), Forcing of low-frequency ocean variability in the Northeast Pacific, Journal of Climate, 22(5), 1255-1276. Chiba, S., K. Tadokoro, H. Sugisaki, and T. Saino (2006), Effects of decadal climate change on zooplankton over the last 50 years in the western subarctic North Pacific, Global Change Biol, 12(5), 907-920. Chiba, S., M. N. Aita, K. Tadokoro, T. Saino, H. Sugisaki, and K. Nakata (2008), From climate regime shifts to lower-trophic level phenology: Synthesis of recent progress in retrospective studies of the western North Pacific, Progress in Oceanography, 77(2-3), 112-126. China-SNAP (2011), The Second National Chinese Assessment Report on Climate Change Chinese Government; Beijing. Christian, J. R., and R. Murtugudde (2003), Tropical Atlantic variability in a coupled physical-biogeochemical ocean model, Deep-Sea Res Pt Ii, 50(22-26), 2947-2969. Church, J. A., N. J. White, and J. R. Hunter (2006), Sea-level rise at tropical Pacific and Indian Ocean islands, Global and Planetary Change, 53(3), 155-168. Church, J. A., N. J. White, R. Coleman, K. Lambeck, and J. X. Mitrovica (2004), Estimates of the regional distribution of sea level rise over the 1950-2000 period, Journal of Climate, 17(13), 2609-2625. CIESM (2008), Climate warming and related changes in Mediterranean marine biota. No. 35 in CIESM Workshop Monographs, edited by F. Briand, p. 152, Monaco. Cinner, J. E., T. R. McClanahan, T. M. Daw, N. A. Graham, J. Maina, S. K. Wilson, and T. P. Hughes (2009), Linking social and ecological systems to sustain coral reef fisheries, Curr Biol, 19(3), 206-212. Clark, C. O., J. E. Cole, and P. J. Webster (2000), Indian Ocean SST and Indian summer rainfall: Predictive relationships and their decadal variability, Journal of Climate, 13(14), 2503-2519. Clifton, J. (2009), Science, funding and participation: key issues for marine protected area networks and the Coral Triangle Initiative, Environmental Conservation, 36(2), 91-96. Cocco, V., F. Joos, M. Steinacher, T. L. Frölicher, L. Bopp, J. Dunne, M. Gehlen, C. Heinze, J. Orr, and A. Oschlies (2013), Oxygen and indicators of stress for marine life in multi-model global warming projections, Biogeosciences, 10(3), 1849-1868. Cochrane, K. L., N. L. Andrew, and A. M. Parma (2011), Primary fisheries management: a minimum requirement for provision of sustainable human benefits in small-­ scale fisheries, Fish Fish, 12(3), 275-288. Cockcroft, A. C., D. van Zyl, and L. Hutchings (2008), Large-scale changes in the spatial distribution of South African West Coast rock lobsters: an overview, African Journal of Marine Science, 30(1), 149-159. Collie, J. S., A. D. Wood, and H. P. Jeffries (2008), Long-term shifts in the species composition of a coastal fish community, Canadian Journal of Fisheries and Aquatic Sciences, 65(7), 1352-1365. Collins, M., S. I. An, W. Cai, A. Ganachaud, E. Guilyardi, F. F. Jin, M. Jochum, M. Lengaigne, S. Power, and A. Timmermann (2010), The impact of global warming on the tropical Pacific Ocean and El Nino, Nature Geoscience, 3(6), 391-397. Colwell, R. R. (1996), Global climate and infectious disease: the cholera paradigm, Science, 274(5295), 2025-2031. Coma, R., M. Ribes, E. Serrano, E. Jiménez, J. Salat, and J. Pascual (2009), Global warming-enhanced stratification and mass mortality events in the Mediterranean, Proceedings of the National Academy of Sciences of the United States of America, 106(15), 6176-6181. Conley, D. J., et al. (2009), Hypoxia-related processes in the Baltic Sea, Environmental Science & Technology, 43(10), 3412-3420. Conversi, A., S. Fonda Umani, T. Peluso, J. C. Molinero, A. Santojanni, and M. Edwards (2010), The Mediterranean Sea regime shift at the end of the 1980s, and intriguing parallelisms with other European basins, Plos One, 5(5), e10633. Cooley, S. R., and S. C. Doney (2009), Anticipating ocean acidification's economic consequences for commercial fisheries, Environ Res Lett, 4(2), 024007. Costello, A., et al. (2009), Managing the health effects of climate change: Lancet and University College London Institute for Global Health Commission, Lancet, 373(9676), 1693-1733. Crain, C. M., K. Kroeker, and B. S. Halpern (2008), Interactive and cumulative effects of multiple human stressors in marine systems, Ecology Letters, 11(12), 1304-1315. Crawford, R. J. M., A. J. Tree, P. A. Whittington, J. Visagie, L. Upfold, K. J. Roxburg, A. P. Martin, and B. M. Dyer (2008), Recent distributional changes of seabirds in South Africa: is climate having an impact?, African Journal of Marine Science, 30(1), 189-193. Subject to Final Copyedit 84 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Crisci, C., N. Bensoussan, J. C. Romano, and J. Garrabou (2011), Temperature anomalies and mortality events in marine communities: insights on factors behind differential mortality impacts in the NW Mediterranean, Plos One, 6(9), e23814. Cruz, A. M., and E. Krausmann (2008), Damage to offshore oil and gas facilities following hurricanes Katrina and Rita: An overview, J Loss Prevent Proc, 21(6), 620-626. Cubillos, J. C., S. W. Wright, G. Nash, M. F. de Salas, B. Griffiths, B. Tilbrook, A. Poisson, and G. M. Hallegraeff (2007), Calcification morphotypes of the coccolithophorid Emiliania huxleyi in the Southern Ocean: changes in 2001 to 2006 compared to historical data, Marine Ecology Progress Series, 348, 47-54. Cury, P., and L. Shannon (2004), Regime shifts in upwelling ecosystems: observed changes and possible mechanisms in the northern and southern Benguela, Progress in Oceanography, 60(2-4), 223-243. Cushing, D. H. (1990), Plankton production and year-class strength in fish populations - an update of the match mismatch hypothesis, Advances in Marine Biology, 26, 249-293. Dalpadado, P., R. B. Ingvaldsen, L. C. Stige, B. Bogstad, T. Knutsen, G. Ottersen, and B. Ellertsen (2012), Climate effects on Barents Sea ecosystem dynamics, ICES Journal of Marine Science, 69(7), 1303-1316. Danovaro, R., S. Fonda Umani, and A. Pusceddu (2009), Climate change and the potential spreading of marine mucilage and microbial pathogens in the Mediterranean Sea, Plos One, 4(9), e7006. Danovaro, R., et al. (2010), Deep-sea biodiversity in the Mediterranean Sea: The known, the unknown, and the unknowable, Plos One, 5(8), e11832. De ath, G., K. E. Fabricius, H. Sweatman, and M. Puotinen (2012), The 27-year decline of coral cover on the Great Barrier Reef and its causes, Proceedings of the National Academy of Sciences of the United States of America, 109(44), 17995-17999. Della Patrona, L., B. Beliaeff, and T. Pickering (2011), Mitigation of sea level rise effects by addition of sediment to shrimp ponds, Aquaculture Environment Interactions, 2, 27-38. Demarcq, H. (2009), Trends in primary production, sea surface temperature and wind in upwelling systems (1998 2007), Progress in Oceanography, 83(1), 376-385. Deser, C., A. S. Phillips, and J. W. Hurrell (2004), Pacific interdecadal climate variability: Linkages between the tropics and the North Pacific during boreal winter since 1900, Journal of Climate, 17(16), 3109-3124. Di Lorenzo, E., et al. (2008), North Pacific Gyre Oscillation links ocean climate and ecosystem change, Geophysical Research Letters, 35(8), L08607. Diaz-Almela, E., N. Marba, and C. M. Duarte (2007), Consequences of Mediterranean warming events in seagrass (Posidonia oceanica) flowering records, Global Change Biol, 13(1), 224-235. Dixson, D. L., P. L. Munday, and G. P. Jones (2010), Ocean acidification disrupts the innate ability of fish to detect predator olfactory cues, Ecology Letters, 13(1), 68-75. Dixson, D. L., G. P. Jones, P. L. Munday, S. Planes, M. S. Pratchett, M. Srinivasan, C. Syms, and S. R. Thorrold (2008), Coral reef fish smell leaves to find island homes, Proceedings of the Royal Society B: Biological Sciences, 275(1653), 2831-2839. Doney, S. C., V. J. Fabry, R. A. Feely, and J. A. Kleypas (2009), Ocean acidification: the other CO2 problem, Annual Review of Marine Science, 1, 169-192. Donner, S. D., C. J. Kucharik, and J. A. Foley (2004), Impact of changing land use practices on nitrate export by the Mississippi River, Global Biogeochemical Cycles, 18(1), GB1028. Donner, S. D., T. R. Knutson, and M. Oppenheimer (2007), Model-based assessment of the role of human-induced climate change in the 2005 Caribbean coral bleaching event, Proceedings of the National Academy of Sciences of the United States of America, 104(13), 5483-5488. Donner, S. D., T. Kirata, and C. Vieux (2010), Recovery from the 2004 coral bleaching event in the Gilbert Islands, Kiribati, Atoll Research Bulletin, 587, 1-27. Donner, S. D., W. J. Skirving, C. M. Little, M. Oppenheimer, and O. Hoegh-Guldberg (2005), Global assessment of coral bleaching and required rates of adaptation under climate change, Global Change Biol, 11(12), 2251-2265. Dove, S., D. I. Kline, O. Pantos, F. E. Angly, G. W. Tyson, and O. Hoegh-Guldberg (2013), Future reef decalcification under a business-as-usual CO2 emission scenario., Proceedings of the National Academy of Sciences of the United States of America, 110(38), 15342-15347. Drinkwater, K. F. (2006), The regime shift of the 1920s and 1930s in the North Atlantic, Progress in Oceanography, 68(2-4), 134-151. Duarte, C. M., J. Middelburg, and N. Caraco (2005), Major role of marine vegetation on the oceanic carbon cycle, Biogeosciences, 2(1), 1-8. Subject to Final Copyedit 85 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Dulvy, N. K., S. I. Rogers, S. Jennings, V. Stelzenmüller, S. R. Dye, and H. R. Skjoldal (2008), Climate change and deepening of the North Sea fish assemblage: a biotic indicator of warming seas, Journal of Applied Ecology, 45(4), 1029-1039. Durack, P. J., and S. E. Wijffels (2010), Fifty-year trends in global ocean salinities and their relationship to broad- scale warming, Journal of Climate, 23(16), 4342-4362. Durack, P. J., S. E. Wijffels, and R. J. Matear (2012), Ocean salinities reveal strong global water cycle intensification during 1950 to 2000, Science, 336(6080), 455-458. Eakin, C. M., et al. (2010), Caribbean corals in crisis: record thermal stress, bleaching, and mortality in 2005, Plos One, 5(11), e13969. Edgar, G. J., et al. (2010), El Nino, grazers and fisheries interact to greatly elevate extinction risk for Galapagos marine species, Global Change Biol, 16(10), 2876-2890. Edwards, M. (2004), Continuous plankton records: Plankton atlas of the North Atlantic Ocean 1958-1999: Preface, Marine Ecology Progress Series, Supplement, 1-2. Edwards, M., and A. J. Richardson (2004), Impact of climate change on marine pelagic phenology and trophic mismatch, Nature, 430(7002), 881-884. Edwards, M., G. Beaugrand, G. C. Hays, J. A. Koslow, and A. J. Richardson (2010), Multi-decadal oceanic ecological datasets and their application in marine policy and management, Trends Ecol Evol, 25(10), 602-610. Elferink, A. G. O. (2012), Governance principles for areas beyond national jurisdiction, The International Journal of Marine and Coastal Law, 27, 205-259. Ellingsen, I. H., P. Dalpadado, D. Slagstad, and H. Loeng (2008), Impact of climatic change on the biological production in the Barents Sea, Climatic Change, 87(1), 155-175. Emanuel, K. (2005), Increasing destructiveness of tropical cyclones over the past 30 years, Nature, 436(7051), 686- 688. Eriksen, E., R. Ingvaldsen, J. E. Stiansen, and G. O. Johansen (2012), Thermal habitat for 0-group fish in the Barents Sea; how climate variability impacts their density, length, and geographic distribution, ICES Journal of Marine Science, 69(5), 870-879. Fabricius, K. E., C. Langdon, S. Uthicke, C. Humphrey, S. Noonan, G. De ath, R. Okazaki, N. Muehllehner, M. S. Glas, and J. M. Lough (2011), Losers and winners in coral reefs acclimatized to elevated carbon dioxide concentrations, Nature Climate Change, 1(3), 165-169. Falkowski, P. G., et al. (2011), Ocean deoxygenation: Past, present, and future, Eos 92(46), 409-410. FAO (2012), State of the World Fisheries and Aquaculture 2012, 197 pp., Food and Agriculture Organization of the United Nations, Rome. Feely, R. A., S. C. Doney, and S. R. Cooley (2009), Ocean acidification: Present conditions and future changes in a high-CO2 world, Oceanography, 22(4), 36-47. Feely, R. A., C. L. Sabine, J. M. Hernandez-Ayon, D. Ianson, and B. Hales (2008), Evidence for upwelling of corrosive "acidified" water onto the continental shelf, Science, 320(5882), 1490-1492. Feely, R. A., C. L. Sabine, K. Lee, W. Berelson, J. Kleypas, V. J. Fabry, and F. J. Millero (2004), Impact of anthropogenic CO2 on the CaCO3 system in the oceans, Science, 305(5682), 362-366. Feely, R. A., C. L. Sabine, R. H. Byrne, F. J. Millero, A. G. Dickson, R. Wanninkhof, A. Murata, L. A. Miller, and D. Greeley (2012), Decadal changes in the aragonite and calcite saturation state of the Pacific Ocean, Global Biogeochemical Cycles, 26(3), GB3001. Firth, L. B., A. M. Knights, and S. S. Bell (2011), Air temperature and winter mortality: Implications for the persistence of the invasive mussel, Perna viridis in the intertidal zone of the south-eastern United States, J Exp Mar Biol Ecol, 400(1-2), 250-256. Fonselius, S., and J. Valderrama (2003), One hundred years of hydrographic measurements in the Baltic Sea, Journal of Sea Research, 49(4), 229-241. Form, A. U., and U. Riebesell (2012), Acclimation to ocean acidification during long-­ term CO2 exposure in the cold-­ water coral Lophelia pertusa, Global Change Biol, 18(3), 843-853. Friedrich, T., et al. (2012), Detecting regional anthropogenic trends in ocean acidification against natural variability, Nature Climate Change, 2(3), 167-171. Frieler, K., M. Meinshausen, A. Golly, M. Mengel, K. Lebek, S. D. Donner, and O. Hoegh-Guldberg (2012), Limiting global warming to 2 °C is unlikely to save most coral reefs, Nature Climate Change, 3, 165-170. Frontiers-Economics-Ltd (2013), Economics of Climate Resilience, Natural Enviornment Theme: Sea fish CA 0401. Report prepared for DEFRA and the devolved administrations.Rep., DEFRA. Subject to Final Copyedit 86 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Fu, B. J., C. H. Su, Y. P. Wei, I. R. Willett, Y. H. Lü, and G. H. Liu (2011), Double counting in ecosystem services valuation: causes and countermeasures, Ecological research, 26(1), 1-14. Galil, B. S. (2008), Alien species in the Mediterranean Sea which, when, where, why?, Hydrobiologia, 606(1), 105-116. Galil, B. S. (2011), The alien crustaceans in the Mediterranean Sea: An historical review, in Invad Nat Spring Ser, edited by B. S. Galil, P. F. Clark and J. T. Carlton, pp. 377-401, Springer Netherlands. Ganachaud, A., et al. (2011), Observed and expected changes to the tropical Pacific Ocean, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. J. Hobday, pp. 101-187, Secretariat of the Pacific Community, Noumea, New Caledonia. Gardner, T. A., I. M. Côté, J. A. Gill, A. Grant, and A. R. Watkinson (2003), Long-term region-wide declines in Caribbean corals, Science, 301(5635), 958-960. Garrabou, J., et al. (2009), Mass mortality in Northwestern Mediterranean rocky benthic communities: effects of the 2003 heat wave, Global Change Biol, 15(5), 1090-1103. Garreaud, R. D., and M. Falvey (2009), The coastal winds off western subtropical South America in future climate scenarios, Int J Climatol, 29(4), 543-554. Gattuso, J.-P., J. Bijma, M. Gehlen, U. Riebesell, and C. Turley (2011), Ocean acidification: knowns, unknowns, and perspectives, in Ocean Acidification, edited by J.-P. Gattuso and L. Hansson, pp. 291-311, Oxford University Press, Oxford, UK. Genner, M. J., D. W. Sims, A. J. Southward, G. C. Budd, P. Masterson, M. McHugh, P. Rendle, E. J. Southall, V. J. Wearmouth, and S. J. Hawkins (2010), Body size-­ dependent responses of a marine fish assemblage to climate change and fishing over a century-­ long scale, Global Change Biol, 16(2), 517-527. Geresi, E., R. Chapman, T. McGee, and B. Woolsey (2009), Monitoring seafloor instability caused by the presence of gas hydrate using ocean acoustical and geophysical techniques in the northern Gulf of Mexico, in Natural Gas Hydrates: Energy resource potential and associated geologic hazards, edited, The American Association of Petroleum Geologists Gillett, N. P., and P. A. Stott (2009), Attribution of anthropogenic influence on seasonal sea level pressure, Geophysical Research Letters, 36(23), L23709. Gillett, N. P., F. W. Zwiers, A. J. Weaver, and P. A. Stott (2003), Detection of human influence on sea-level pressure, Nature, 422(6929), 292-294. Gilman, E., J. Ellison, and R. Coleman (2007), Assessment of mangrove response to projected relative sea-level rise and recent historical reconstruction of shoreline position, Environmental Monitoring and Assessment, 124(1-3), 105-130. Gleckler, P. J., B. Santer, C. Domingues, D. Pierce, T. Barnett, J. Church, K. Taylor, K. AchutaRao, T. Boyer, and M. Ishii (2012), Human-induced global ocean warming on multidecadal timescales, Nature Climate Change, 2, 524-529. Gledhill, D. K., R. Wanninkhof, F. J. Millero, and M. Eakin (2008), Ocean acidification of the Greater Caribbean Region 1996 2006, Journal of Geophysical Research, 113(C10), C10031. Glynn, P. W. (1984), Widespread coral mortality and the 1982-83 El Nino warming event, Environmental Conservation, 11(2), 133-146. Glynn, P. W. (2001), Eastern Pacific coral reef ecosystems, in Coastal marine ecosystems of Latin America. Ecological Studies 144, edited by U. Seeliger and B. Kjerfve, pp. 281-305, Springer-Verlag, Berlin & Heidelberg. Glynn, P. W. (2011), In tandem reef coral and cryptic metazoan declines and extinctions, B Mar Sci, 87(4), 767-794. Gómez-Gesteira, M., M. De Castro, I. Álvarez, M. N. Lorenzo, J. L. G. Gesteira, and A. J. C. Crespo (2008), Spatio- temporal upwelling trends along the Canary Upwelling System (1967 2006), Annals of the New York Academy of Sciences, 1146(1), 320-337. González-Taboada, F., and R. Anadón (2012), Patterns of change in sea surface temperature in the North Atlantic during the last three decades: beyond mean trends, Climatic Change, 115, 491-431. Goreau, T., T. McClanahan, R. Hayes, and A. Strong (2000), Conservation of coral reefs after the 1998 global bleaching event, Conservation Biology, 14(1), 5-15. GPO (2012), Framework Document, Global Partnership for OceansRep., World Bank, Washington DC. Graham, N. A., S. K. Wilson, S. Jennings, N. V. Polunin, J. Robinson, J. P. Bijoux, and T. M. Daw (2007), Lag effects in the impacts of mass coral bleaching on coral reef fish, fisheries, and ecosystems, Conservation Biology, 21(5), 1291-1300. Subject to Final Copyedit 87 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Graham, N. A., et al. (2008), Climate warming, marine protected areas and the ocean-scale integrity of coral reef ecosystems, Plos One, 3(8), e3039. Grantham, B. A., F. Chan, K. J. Nielsen, D. S. Fox, J. A. Barth, A. Huyer, J. Lubchenco, and B. A. Menge (2004), Upwelling-driven nearshore hypoxia signals ecosystem and oceanographic changes in the northeast Pacific, Nature, 429(6993), 749-754. Green, D., L. Alexander, K. Mclnnes, J. Church, N. Nicholls, and N. White (2010), An assessment of climate change impacts and adaptation for the Torres Strait Islands, Australia, Climatic Change, 102(3-4), 405-433. Gregg, W. W., N. W. Casey, and C. R. McClain (2005), Recent trends in global ocean chlorophyll, Geophysical Research Letters, 32(3), L03606. Griffith, G. P., E. A. Fulton, R. Gorton, and A. J. Richardson (2012), Predicting interactions among fishing, ocean warming, and ocean acidification in a marine system with whole-­ ecosystem models, Conservation Biology, 26(6), 1145-1152. Gruber, N. (2011), Warming up, turning sour, losing breath: ocean biogeochemistry under global change, Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 369(1943), 1980-1996. Gruber, N., C. Hauri, Z. Lachkar, D. Loher, T. L. Frölicher, and G. K. Plattner (2012), Rapid progression of ocean acidification in the California Current System, Science, 337(6091), 220-223. Guilyardi, E., H. Bellenger, M. Collins, S. Ferrett, W. Cai, and A. Wittenberg (2012), A first look at ENSO in CMIP5, CLIVAR Exchanges No.58, 17(1), 29-32. Guinotte, J. M., J. Orr, S. Cairns, A. Freiwald, L. Morgan, and R. George (2006), Will human-induced changes in seawater chemistry alter the distribution of deep-sea scleractinian corals?, Front Ecol Environ, 4(3), 141-146. Gutierrez, D., et al. (2011), Coastal cooling and increased productivity in the main upwelling zone off Peru since the mid-twentieth century, Geophysical Research Letters, 38, L07603. Hall-Spencer, J. M., R. Rodolfo-Metalpa, S. Martin, E. Ransome, M. Fine, S. M. Turner, S. J. Rowley, D. Tedesco, and M.-C. Buia (2008), Volcanic carbon dioxide vents show ecosystem effects of ocean acidification, Nature, 454(7200), 96-99. Halpern, B. S., S. E. Lester, and K. L. McLeod (2010), Placing marine protected areas onto the ecosystem-based management seascape, Proceedings of the National Academy of Sciences of the United States of America, 107(43), 18312-18317. Halpern, B. S., K. L. McLeod, A. A. Rosenberg, and L. B. Crowder (2008), Managing for cumulative impacts in ecosystem-based management through ocean zoning, Ocean & Coastal Management, 51(3), 203-211. Halpern, B. S., C. M. Ebert, C. V. Kappel, E. M. P. Madin, F. Micheli, M. Perry, K. A. Selkoe, and S. Walbridge (2009), Global priority areas for incorporating land-sea connections in marine conservation, Conservation Letters, 2(4), 189-196. Hänninen, J., and I. Vuorinen (2011), Time-varying parameter analysis of the Baltic Sea freshwater runoffs, Environ Model Assess, 16(1), 53-60. Hare, S. R., and N. J. Mantua (2000), Empirical evidence for North Pacific regime shifts in 1977 and 1989, Progress in Oceanography, 47(2-4), 103-145. Harley, C. D. (2011), Climate change, keystone predation, and biodiversity loss, Science, 334(6059), 1124-1127. Harvell, C. D., C. E. Mitchell, J. R. Ward, S. Altizer, A. P. Dobson, R. S. Ostfeld, and M. D. Samuel (2002), Ecology - Climate warming and disease risks for terrestrial and marine biota, Science, 296(5576), 2158-2162. Harvell, C. D., et al. (1999), Emerging marine diseases-climate links and anthropogenic factors, Science, 285(5433), 1505-1510. Harvell, D., et al. (2004), The rising tide of ocean diseases: unsolved problems and research priorities, Front Ecol Environ, 2(7), 375-382. Hashioka, T., T. T. Sakamoto, and Y. Yamanaka (2009), Potential impact of global warming on North Pacific spring blooms projected by an eddy-permitting 3-D ocean ecosystem model, Geophysical Research Letters, 36, L20604. Hátún, H., M. Payne, G. Beaugrand, P. Reid, A. Sand, H. Drange, B. Hansen, J. Jacobsen, and D. Bloch (2009), Large bio-geographical shifts in the north-eastern Atlantic Ocean: From the subpolar gyre, via plankton, to blue whiting and pilot whales, Progress in Oceanography, 80(3-4), 149-162. Hazen, E. L., S. Jorgensen, R. R. Rykaczewski, S. J. Bograd, D. G. Foley, I. D. Jonsen, S. A. Shaffer, J. P. Dunne, D. P. Costa, and L. B. Crowder (2012), Predicted habitat shifts of Pacific top predators in a changing climate, Nature Climate Change, 3, 234-238. Subject to Final Copyedit 88 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Head, E. J. H., and P. Pepin (2010), Monitoring changes in phytoplankton abundance and composition in the Northwest Atlantic: a comparison of results obtained by continuous plankton recorder sampling and colour satellite imagery, J Plankton Res, 32(12), 1649-1660. Healy, T., and K. Harada (1991), Definition and physical characteristics of the world's enclosed coastal seas, Mar Pollut Bull, 23, 639-644. Hegerl, G. C., O. Hoegh-Guldberg, G. Casassa, M. P. Hoerling, R. S. Kovats, C. Parmesan, D. W. Pierce, and P. A. Stott (2010), Good practice guidance paper on detection and attribution related to anthropogenic climate change, in Meeting Report of the Intergovernmental Panel on Climate Change Expert Meeting on Detection and Attribution of Anthropogenic Climate Change, edited by T. F. Stocker, C. B. Field, D. Qin, V. Barros, G.-K. Plattner, M. Tignor, P. M. Midgley and K. L. Ebi, p. 8, University of Bern, Bern, Switzerland. Hegerl, G. C., F. W. Zwiers, P. Braconnot, N. P. Gillett, Y. Luo, J. A. Marengo Orsini, N. Nicholls, J. T. Penner, and P. Stott (2007), Understanding and Attributing Climate Change, in Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, edited by S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, M. Tignor and H. L. Miller, Cambridge University Press, Cambridge, UK and New York, NY, USA. Heil, C. A., P. M. Glibert, M. A. Al-Sarawi, M. Faraj, M. Behbehani, and M. Husain (2001), First record of a fish- killing Gymnodinium sp. bloom in Kuwait Bay, Arabian Sea: chronology and potential causes, Marine Ecology Progress Series, 214, 15-23. Heiss, G. A. (1996), Annual band width variation in Porites sp. from Aqaba, Gulf of Aqaba, Red Sea, B Mar Sci, 59(2), 393-403. Heitmann, N., and S. Khalilian (2010), Accounting for CO2 emissions from international shipping: burden sharing under different UNFCCC allocation options and regime scenarios. Helly, J. J., and L. A. Levin (2004), Global distribution of naturally occurring marine hypoxia on continental margins, Deep-Sea Research Part I-Oceanographic Research Papers, 51(9), 1159-1168. Helm, K. P., N. L. Bindoff, and J. A. Church (2011a), Observed decreases in oxygen content of the global ocean, Geophysical Research Letters, 38(23). Helm, K. P., N. L. Bindoff, and J. A. Church (2011b), Observed decreases in oxygen content of the global ocean, Geophysical Research Letters, 38(L23602), 1-6. Henriksen, P. (2009), Long-term changes in phytoplankton in the Kattegat, the Belt Sea, the Sound and the western Baltic Sea, Journal of Sea Research, 61(1-2), 114-123. Henson, S. A., J. L. Sarmiento, J. P. Dunne, L. Bopp, I. D. Lima, S. C. Doney, J. John, and C. Beaulieu (2010), Detection of anthropogenic climate change in satellite records of ocean chlorophyll and productivity, Biogeosciences, 7, 621-640. Hermansen, O., and K. Heen (2012), Norwegian salmonid farming and global warming: socioeconomic impacts, Aquaculture Economics & Management, 16(3), 220-221. Hermant, M., J. Lobry, S. Bonhommeau, J. C. Poulard, and O. Le Pape (2010), Impact of warming on abundance and occurrence of flatfish populations in the Bay of Biscay (France), Journal of Sea Research, 64(1-2), 45-53. Heymans, J. J., L. J. Shannon, and A. Jarre (2004), Changes in the northern Benguela ecosystem over three decades: 1970s, 1980s, and 1990s, Ecological Modelling, 172(2-4), 175-195. Hickey, B. M. (1979), The California Current System-hypotheses and facts, Progress in Oceanography, 8(4), 191- 279. Hiddink, J. G., and R. ter Hofstede (2008), Climate induced increases in species richness of marine fishes, Global Change Biol, 14(3), 453-460. Hilborn, R. (2007), Defining success in fisheries and conflicts in objectives, Mar Policy, 31(2), 153-158. Hill, K. L., S. R. Rintoul, R. Coleman, and K. R. Ridgway (2008), Wind forced low frequency variability of the East Australia Current, Geophysical Research Letters, 35(8), L08602. Hilmi, N., D. Allemand, S. Dupont, A. Safa, G. Haraldsson, P. A. L. D. Nunes, C. Moore, C. Hattam, S. Reynaud, and J. M. Hall-Spencer (2013), Towards improved socio-economic assessments of ocean acidification s impacts, Mar. Biol., 160, 1773-1787. Hitchcock, G. L., P. Lane, S. Smith, J. G. Luo, and P. B. Ortner (2002), Zooplankton spatial distributions in coastal waters of the northern Arabian Sea, August, 1995, Deep-Sea Res Pt Ii, 49(12), 2403-2423. Hjort, J. (1914), Fluctuations in the great fisheries of northern Europe, Rapports et proces-verbaux des reunions/Conseil permanent international pour l'exploration de la mer, 20, 1-228. Subject to Final Copyedit 89 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Hoegh-Guldberg, O. (1999), Climate change, coral bleaching and the future of the world's coral reefs, Marine and Freshwater Research, 50(8), 839-866. Hoegh-Guldberg, O. (2011), Coral reef ecosystems and anthropogenic climate change, Regional Environmental Change, 11, S215-S227. Hoegh-Guldberg, O. (2012), The adaptation of coral reefs to climate change: Is the Red Queen being outpaced?, Scientia Marina, 76(2), 403-408. Hoegh-Guldberg, O., and B. Salvat (1995), Periodic mass-bleaching and elevated sea temperatures - bleaching of outer reef slope communities in Moorea, French Polynesia, Marine Ecology Progress Series, 121(1-3), 181- 190. Hoegh-Guldberg, O., and J. F. Bruno (2010), The impact of climate change on the world's marine ecosystems, Science, 328(5985), 1523-1528. Hoegh-Guldberg, O., L. Hughes, S. McIntyre, D. B. Lindenmayer, C. Parmesan, H. P. Possingham, and C. D. Thomas (2008), Assisted colonization and rapid climate change, Science, 321(5887), 345-346. Hoegh-Guldberg, O., S. Andréfouët, K. E. Fabricius, G. Diaz-Pulido, J. M. Lough, P. A. Marshall, and M. S. Pratchett (2011a), Vulnerability of coral reefs in the tropical Pacific to climate change, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. L. Hobday, pp. 251-296, Secretariat of the Pacific Community, Noumea, New Caledonia. Hoegh-Guldberg, O., G. Hegerl, T. Root, F. Zwiers, P. Stott, D. Pierce, and M. Allen (2011b), Correspondence: Difficult but not impossible, Nature Climate Change, 1(2), 72-72. Hoegh-Guldberg, O., et al. (2007), Coral reefs under rapid climate change and ocean acidification, Science, 318(5857), 1737-1742. Hoegh-Guldberg, O., et al. (2009), The Coral Triangle and Climate Change: ecosystems, people and societies at riskRep., 276 pp, WWF Australia, Brisbane. Hofmann, A. F., E. T. Peltzer, and P. G. Brewer (2013), Kinetic bottlenecks to respiratory exchange rates for deep- sea-Part I: Oxygen, Biogeosciences Discussions, 10, 5049-5060. Hönisch, B., et al. (2012), The geological record of ocean acidification, Science, 335(6072), 1058-1063. Howard, J. A., A. Jarre, A. Clark, and C. Moloney (2007), Application of the sequential t-test algorithm for analysing regime shifts to the southern Benguela ecosystem, African Journal of Marine Science, 29(3), 437- 451. Howell, E. A., C. C. C. Wabnitz, J. P. Dunne, and J. J. Polovina (2013), Climate-induced primary productivity change and fishing impacts on the Central North Pacific ecosystem and Hawaii-based pelagic longline fishery, Climatic Change, 119(1), 79-93. Hughes, T. P. (1994), Catastrophes, phase shifts, and large-scale degradation of a Caribbean coral reef, Science, 265(5178), 1547-1551. Hughes, T. P., D. R. Bellwood, A. H. Baird, J. Brodie, J. F. Bruno, and J. M. Pandolfi (2011), Shifting base-lines, declining coral cover, and the erosion of reef resilience: comment on Sweatman et al. (2011), Coral Reefs, 30(3), 653-660. Hughes, T. P., et al. (2003), Climate change, human impacts, and the resilience of coral reefs, Science, 301(5635), 929-933. Hunt, G. L., P. Stabeno, G. Walters, E. Sinclair, R. D. Brodeur, J. M. Napp, and N. A. Bond (2002), Climate change and control of the southeastern Bering Sea pelagic ecosystem, Deep-Sea Res Pt Ii, 49(26), 5821-5853. Hunt, G. L., et al. (2010), Status and trends of the Bering Sea region, 2003-2008, in Marine Ecosystems of the North Pacific Ocean, 2003-2008. PICES Special Publication 4, 393 p, edited by S. M. McKinnell and M. J. Dagg, pp. 196-267. Huse, G., and I. Ellingsen (2008), Capelin migrations and climate change a modelling analysis, Climatic Change, 87(1), 177-197. Huse, G., J. C. Holst, K. Utne, L. Nttestad, W. Melle, A. Slotte, G. Ottersen, T. Fenchel, and F. Uiblein (2012), Effects of interactions between fish populations on ecosystem dynamics in the Norwegian Sea results of the INFERNO project, Marine Biology Research, 8(5-6), 415-419. Hutchings, L., et al. (2009), The Benguela Current: An ecosystem of four components, Progress in Oceanography, 83(1-4), 15-32. ICES (2012), Report of the ICES Advisory Committee 2012. ICES Advice, 2012. Book 3Rep., 82 pp. Subject to Final Copyedit 90 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Iles, A. C., T. C. Gouhier, B. A. Menge, J. S. Stewart, A. J. Haupt, and M. C. Lynch (2012), Climate-­ driven trends and ecological implications of event-­ scale upwelling in the California Current System, Global Change Biol, 18(2), 783-796. IPCC (2005), IPCC Special Report on Carbon Dioxide Capture and Storage. Prepared by Working Group III of the Intergovernmental Panel on Climate ChangeRep., 442 pp, Intergovernmental Panel on Climate Change Working Group III, Cambridge, United Kingdom and New York, NY, USA. IPCC (2012), Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation, 582 pp, Cambridge University Press, Cambridge, UK, and New York, NY, USA. Irvine, J. R., and M. A. Fukuwaka (2011), Pacific salmon abundance trends and climate change, ICES Journal of Marine Science, 68(6), 1122-1130. Ito, S., H. Sugisaki, A. Tsuda, O. Yamamura, and K. Okuda (2004), Contributions of the VENFISH program: meso-­ zooplankton, Pacific saury (Cololabis saira) and walleye pollock (Theragra chalcogramma) in the northwestern Pacific, Fisheries Oceanography, 13(s1), 1-9. Ito, S., K. A. Rose, A. J. Miller, K. Drinkwater, K. Brander, J. E. Overland, S. Sundby, E. Curchitser, J. W. Hurrell, and Y. Yamanaka (2010), Ocean ecosystem responses to future global change scenarios: a way forward, in Marine Ecosystems and Global Change, edited by M. Barange, J. G. Field, R. H. Harris, E. Hofmann, R. I. Perry and F. Werner, pp. 287-322, Oxford University Press, Oxford, UK. Ittekkot, V., S. Jilan, E. Miles, E. Desa, B. N. Desai, J. T. Everett, J. J. Magnuson, A. Tsyban, and S. Zuta (1996), Oceans, in Climate Change 1995: Impacts, Adaptations, and Mitigation of Climate Change: Scientific- Technical Analyses. Contribution of Working Group II to the Second Assessment Report of the Intergovernmental Panel on Climate Change, edited by R. T. Watson, M. Zinyowera and R. Moss, pp. 267 288, Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA. Iversen, T. (1934), Some observations on cod in northern waters, Report on Norwegian Fishery and Marine Investigations, 4, 1-35. Jenkins, W. J. (2008), The biogeochemical consequences of changing ventilation in the Japan/East Sea, Mar Chem, 108(3-4), 137-147. Jin, M. B., C. Deal, J. Wang, and C. P. McRoy (2009), Response of lower trophic level production to long-term climate change in the southeastern Bering Sea, Journal of Geophysical Research, 114, C04010. Johannesen, E., R. B. Ingvaldsen, B. Bogstad, P. Dalpadado, E. Eriksen, H. Gjsaeter, T. Knutsen, M. Skern- Mauritzen, and J. E. Stiansen (2012), Changes in Barents Sea ecosystem state, 1970 2009: climate fluctuations, human impact, and trophic interactions, ICES Journal of Marine Science, 69(5), 880-889. Jorda, G., D. Gomis, E. Alvarez-Fanjul, and S. Somot (2012), Atmospheric contribution to Mediterranean and nearby Atlantic sea level variability under different climate change scenarios, Global and Planetary Change, 80-81, 198-214. Jung, S. (2008), Spatial variability in long-term changes of climate and oceanographic conditions in Korea, J Environ Biol, 29(4), 519-529. Jung, S., I.-C. Pang, J.-H. Lee, I. Choi, and H. K. Cha (2013), Latitudinal shifts in the distribution of exploited fishes in Korean waters during the last 30 years: a consequence of climate change, Reviews in Fisheries Biology and Fisheries. Jury, M., T. McClanahan, and J. Maina (2010), West Indian Ocean variability and East African fish catch, Mar Environ Res, 70(2), 162-170. Justiæ, D., N. N. Rabalais, and R. E. Turner (1996), Effects of climate change on hypoxia in coastal waters: A doubled CO2 scenario for the northern Gulf of Mexico, Limnol Oceanogr, 41(5), 992-1003. Justiæ, D., V. J. Bierman, D. Scavia, and R. D. Hetland (2007), Forecasting Gulf's hypoxia: The next 50 years?, Estuar Coast, 30(5), 791-801. Kalnay, E., et al. (1996), The NCEP/NCAR 40-Year Reanalysis Project, Bulletin of the American Meteorological Society, 77(3), 437-471. Kane, J. (2007), Zooplankton abundance trends on Georges Bank, 1977-2004, ICES Journal of Marine Science, 64(5), 909-919. Karstensen, J., L. Stramma, and M. Visbeck (2008), Oxygen minimum zones in the eastern tropical Atlantic and Pacific oceans, Progress in Oceanography, 77(4), 331-350. Kaye, S. (2012), Climate change and maritime security, in Climate Change and the Oceans; Gauging the Legal and Policy Currents in the Asia Pacific and Beyond, edited by R. Warner and C. Schofield, pp. 153-166, Edward Elgar Publishing, Cheltenham, UK; Northampton, MA, USA. Subject to Final Copyedit 91 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Keeling, R. E., A. Körtzinger, and N. Gruber (2010), Ocean deoxygenation in a warming world, Annual Review of Marine Science, 2, 199-229. Keister, J. E., E. Di Lorenzo, C. A. Morgan, V. Combes, and W. T. Peterson (2011), Zooplankton species composition is linked to ocean transport in the Northern California Current, Global Change Biol, 17(7), 2498- 2511. Kelly, R. P., M. M. Foley, W. S. Fisher, R. A. Feely, B. S. Halpern, G. G. Waldbusser, and R. R. Caldwell (2011), Mitigating local causes of ocean acidification with existing laws, Science, 332(6033), 1036-1037. Kenchington, R., and R. Warner (2012), Uncertain seas ahead: legal and policy approaches to conserving marine biodiversity in the face of a changing climate, in Climate Change and the Oceans; Gauging the Legal and Policy Currents in the Asia Pacific and Beyond, edited by R. Warner and C. Schofield, pp. 153-166, Edward Elgar Publishing, Cheltenham, UK; Northampton, MA, USA. Kerr, R. A. (2000), A North Atlantic climate pacemaker for the centuries, Science, 288(5473), 1984-1985. Kessler, J. D., et al. (2011), A persistent oxygen anomaly reveals the fate of spilled methane in the deep Gulf of Mexico, Science, 331(6015), 312-315. Key, R. M., A. Kozyr, C. L. Sabine, K. Lee, R. Wanninkhof, J. L. Bullister, R. A. Feely, F. J. Millero, C. W. Mordy, and T. H. Peng (2004), A global ocean carbon climatology: Results from Global Data Analysis Project (GLODAP), Global Biogeochemical Cycles, 18(4), GB4031. Kideys, A. E. (2002), Fall and rise of the Black Sea ecosystem, Science, 297(5586), 1482-1484. King, J. R., V. N. Agostini, C. J. Harvey, G. A. McFarlane, M. G. G. Foreman, J. E. Overland, E. Di Lorenzo, N. A. Bond, and K. Y. Aydin (2011), Climate forcing and the California Current ecosystem, ICES Journal of Marine Science, 68(6), 1199-1216. Kirby, R. R., and G. Beaugrand (2009), Trophic amplification of climate warming, Proceedings of the Royal Society B: Biological Sciences, 276(1676), 4095-4103. Kirkman, S. P., D. Yemane, J. Kathena, S. K. Mafwila, S. E. Nsiangango, T. Samaai, B. Axelsen, and L. Singh (2013), Identifying and characterizing demersal fish biodiversity hotspots in the Benguela Current Large Marine Ecosystem: relevance in the light of global changes, ICES Journal of Marine Science: Journal du Conseil. Kleypas, J. A., R. W. Buddemeier, D. Archer, J.-P. Gattuso, C. Langdon, and B. N. Opdyke (1999), Geochemical consequences of increased atmospheric carbon dioxide on coral reefs, Science, 284(5411), 118-120. Knoll, A. H., and W. W. Fischer (2011), Skeletons and ocean chemistry: the long view, in Ocean Acidification, edited by J.-P. Gattuso and L. Hansson, pp. 67-82, Oxford University Press, Oxford, UK. Knutson, T. R., J. L. McBride, J. Chan, K. Emanuel, G. Holland, C. Landsea, I. Held, J. P. Kossin, A. Srivastava, and M. Sugi (2010), Tropical cyclones and climate change, Nature Geoscience, 3(3), 157-163. Koslow, J. A., R. Goericke, A. Lara-Lopez, and W. Watson (2011), Impact of declining intermediate-water oxygen on deepwater fishes in the California Current, Marine Ecology Progress Series, 436, 207-218. Krishnan, P., S. D. Roy, G. George, R. C. Srivastava, A. Anand, S. Murugesan, M. Kaliyamoorthy, N. Vikas, and R. Soundararajan (2011), Elevated sea surface temperature during May 2010 induces mass bleaching of corals in the Andaman, Curr Sci India, 100(1), 111-117. Kroeker, K. J., R. L. Kordas, R. Crim, I. E. Hendriks, L. Ramajo, G. S. Singh, C. M. Duarte, and J.-P. Gattuso (2013), Impacts of ocean acidification on marine organisms: quantifying sensitivies and interaction with warming, Global Change Biol, 19, 1884-1896. Kurien, J., and R. Willmann (2009), Special considerations for small-­ scale fisheries management in developing countries, in A Fishery Manager's Guidebook, edited by K. Cochrane and S. M. Garcia, pp. 404-424, FAO and Wiley-Blackwell, Rome, UK and Chichester, UK. Kuzmina, S. I., L. Bengtsson, O. M. Johannessen, H. Drange, L. P. Bobylev, and M. W. Miles (2005), The North Atlantic Oscillation and greenhouse-gas forcing, Geophysical Research Letters, 32(4), L04703. Laffoley, D., and G. D. Grimsditch (2009), The management of natural coastal carbon sinks, Iucn. Last, P. R., W. T. White, D. C. Gledhill, A. J. Hobday, R. Brown, G. J. Edgar, and G. Pecl (2011), Long-term shifts in abundance and distribution of a temperate fish fauna: a response to climate change and fishing practices, Global Ecology and Biogeography, 20(1), 58-72. Laws, E. A., P. G. Falkowski, W. O. Smith, H. Ducklow, and J. J. McCarthy (2000), Temperature effects on export production in the open ocean, Global Biogeochemical Cycles, 14(4), 1231-1246. Le Borgne, R., R. T. Barber, T. Delcroix, H. Y. Inoue, D. J. Mackey, and M. Rodier (2002), Pacific warm pool and divergence: temporal and zonal variations on the equator and their effects on the biological pump, Deep-Sea Res Pt Ii, 49(13-14), 2471-2512. Subject to Final Copyedit 92 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Le Borgne, W. R., V. Allain, S. P. Griffiths, R. J. Matear, A. D. McKinnon, A. J. Richardson, and J. W. Young (2011), Vulnerability of open oceans food webs in the tropical Pacific to climate change, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. J. Hobday, pp. 189-250, Secretariat of the Pacific Community, Noumea, New Caledonia. Leduc, G., C. T. Herbert, T. Blanz, P. Martinez, and R. Schneider (2010), Contrasting evolution of sea surface temperature in the Benguela upwelling system under natural and anthropogenic climate forcings, Geophysical Research Letters, 37(20), L20705. Lee, S.-K., W. Park, E. van Sebille, M. O. Baringer, C. Wang, D. B. Enfield, S. G. Yeager, and B. P. Kirtman (2011), What caused the significant increase in Atlantic Ocean heat content since the mid-20th century?, Geophysical Research Letters, 38(17), L17607. Lee, S. S., J. Y. Lee, B. Wang, K. J. Ha, K. Y. Heo, F. F. Jin, D. M. Straus, and J. Shukla (2012), Interdecadal changes in the storm track activity over the North Pacific and North Atlantic, Clim Dynam, 39(1), 313-327. Lee, T., and M. J. McPhaden (2010), Increasing intensity of El Nino in the central-equatorial Pacific, Geophysical Research Letters, 37, L14603. Lee, Z., S. L. Shang, C. M. Hu, M. Lewis, R. Arnone, Y. H. Li, and B. Lubac (2010), Time series of bio-optical properties in a subtropical gyre: Implications for the evaluation of interannual trends of biogeochemical properties, Journal of Geophysical Research, 115, C09012. Lehodey, P., I. Senina, and R. Murtugudde (2008), A spatial ecosystem and populations dynamics model (SEAPODYM) - Modeling of tuna and tuna-like populations, Progress in Oceanography, 78(4), 304-318. Lehodey, P., R. Murtugudde, and I. Senina (2010a), Bridging the gap from ocean models to population dynamics of large marine predators: A model of mid-trophic functional groups, Progress in Oceanography, 84(1-2), 69-84. Lehodey, P., M. Bertignac, J. Hampton, A. Lewis, and J. Picaut (1997), El Nino Southern Oscillation and tuna in the western Pacific, Nature, 389(6652), 715-718. Lehodey, P., I. Senina, B. Calmettes, J. Hampton, and S. Nicol (2013), Modelling the impact of climate change on Pacific skipjack tuna population and fisheries, Climatic Change, 119(1), 95-109. Lehodey, P., I. Senina, J. Sibert, L. Bopp, B. Calmettes, J. Hampton, and R. Murtugudde (2010b), Preliminary forecasts of Pacific bigeye tuna population trends under the A2 IPCC scenario, Progress in Oceanography, 86, 302-315. Lehodey, P., et al. (2011), Vulnerability of the ocean fisheries in the tropical Pacific to climate change, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change., edited by J. D. Bell, J. E. Johnson and A. L. Hobday, pp. 433-492, Secretariat of the Pacific Community, Noumea, New Caledonia. Lehodey, P., et al. (2006), Climate variability, fish, and fisheries, Journal of Climate, 19(20), 5009-5030. Lejeusne, C., P. Chevaldonné, C. Pergent-Martini, C. F. Boudouresque, and T. Pérez (2010), Climate change effects on a miniature ocean: the highly diverse, highly impacted Mediterranean Sea, Trends Ecol Evol, 25(4), 250-260. Levin, L. A., W. Ekau, A. J. Gooday, F. Jorissen, J. J. Middelburg, S. W. A. Naqvi, C. Neira, N. N. Rabalais, and J. Zhang (2009), Effects of natural and human-induced hypoxia on coastal benthos, Biogeosciences, 6(10), 2063- 2098. Levitus, S., J. Antonov, and T. Boyer (2005), Warming of the world ocean, 1955 2003, Geophysical Research Letters, 32(2), L02604. Levitus, S., J. I. Antonov, T. P. Boyer, R. A. Locarnini, H. E. Garcia, and A. V. Mishonov (2009), Global ocean heat content 1955-2008 in light of recently revealed instrumentation problems, Geophysical Research Letters, 36(7), L07608. Li, L., J. Xu, and R. Cai (2002), Trends of sea level rise in the South China Sea during the 1990s: An altimetry result, Chinese Science Bulletin, 47(7), 82-585. Li, S., K. F. Yu, T. R. Chen, Q. Shi, and H. L. Zhang (2011), Assessment of coral bleaching using symbiotic zooxanthellae density and satellite remote sensing data in the Nansha Islands, South China Sea, Chinese Science Bulletin, 56(10), 1031-1037. Lin, C., X. Ning, J. Su, Y. Lin, and B. Xu (2005), Environmental changes and the responses of the ecosystems of the Yellow Sea during 1976 2000, J Marine Syst, 55(3), 223-234. Lin, G., and Q. Yang (2011), Interdecadal variation of distribution features of phytoplankton in the Taiwan Strait to global climate change, Chinese Journal of Applied & Environmental Biology, 17(5), 615-623. Lindegren, M., C. Möllmann, A. Nielsen, K. Brander, B. R. MacKenzie, and N. C. Stenseth (2010), Ecological forecasting under climate change: the case of Baltic cod, Proceedings of the Royal Society B: Biological Sciences, 277(1691), 2121-2130. Subject to Final Copyedit 93 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Lindley, J. A., and R. R. Kirby (2010), Climate-induced changes in the North Sea Decapoda over the last 60 years, Clim Res, 42(3), 257-264. Lindley, J. A., G. Beaugrand, C. Luczak, J. M. Dewarumez, and R. R. Kirby (2010), Warm-water decapods and the trophic amplification of climate in the North Sea, Biology letters, 6(6), 773-776. Ling, S. D. (2008), Range expansion of a habitat-modifying species leads to loss of taxonomic diversity: a new and impoverished reef state, Oecologia, 156(4), 883-894. Ling, S. D., C. R. Johnson, S. Frusher, and C. K. King (2008), Reproductive potential of a marine ecosystem engineer at the edge of a newly expanded range, Global Change Biol, 14(4), 907-915. Ling, S. D., C. R. Johnson, S. D. Frusher, and K. R. Ridgway (2009), Overfishing reduces resilience of kelp beds to climate-driven catastrophic phase shift, Proceedings of the National Academy of Sciences of the United States of America, 106(52), 22341-22345. Link, J. S., B. Bogstad, H. Sparholt, and G. R. Lilly (2009a), Trophic role of Atlantic cod in the ecosystem, Fish Fish, 10(1), 58-87. Link, J. S., et al. (2009b), A comparison of biological trends from four marine ecosystems: Synchronies, differences, and commonalities, Progress in Oceanography, 81(1-4), 29-46. Litzow, M. A., and L. Ciannelli (2007), Oscillating trophic control induces community reorganization in a marine ecosystem, Ecology Letters, 10(12), 1124-1134. Litzow, M. A., J. D. Urban, and B. J. Laurel (2008), Increased spatial variance accompanies reorganization of two continental shelf ecosystems, Ecol Appl, 18(6), 1331-1337. Liu, C., Y. Du, Q. Zhang, T. Chen, and D. Wang (2007), Responses of oceans to global warming and observation evidence in the South China Sea, Advances in Climate Change Research, 3(1), 8-13. Livingston, P. A., and S. Tjelmeland (2000), Fisheries in boreal ecosystems, ICES Journal of Marine Science, 57(3), 619-627. Llewellyn, L. E. (2010), Revisiting the association between sea surface temperature and the epidemiology of fish poisoning in the South Pacific: Reassessing the link between ciguatera and climate change, Toxicon, 56(5), 691- 697. Lo Iacono, C., M. A. Mateo, E. Gracia, L. Guasch, R. Carbonell, L. Serrano, O. Serrano, and J. Danobeitia (2008), Very high-resolution seismo-acoustic imaging of seagrass meadows (Mediterranean Sea): Implications for carbon sink estimates, Geophysical Research Letters, 35, L18601. Longhurst, A. (1998), Ecological geography of the sea, Academic Press, San Diego, USA. Lotze, H. K., M. Coll, A. M. Magera, C. Ward-Paige, and L. Airoldi (2011), Recovery of marine animal populations and ecosystems, Trends Ecol Evol, 26(11), 595-605. Lough, J. M. (2012), Small change, big difference: Sea surface temperature distributions for tropical coral reef ecosystems, 1950 2011, Journal of Geophysical Research: Oceans (1978 2012), 117(C9), C09018. Lough, J. M., G. A. Meehl, and M. J. Salinger (2011), Observed in projected changes in surface climate of the tropical Pacific, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. L. Hobday, pp. 49-99, Secretariat of the Pacific Community, Noumea, New Caledonia. Lucey, S. M., and J. A. Nye (2010), Shifting species assemblages in the Northeast US Continental Shelf Large Marine Ecosystem, Marine Ecology Progress Series, 415, 23-33. Luchetta, A., C. Cantoni, and G. Catalano (2010), New observations of CO2-induced acidification in the northern Adriatic Sea over the last quarter century, Chem Ecol, 26, 1-17. Ludwig, W., E. Dumont, M. Meybeck, and S. Heussner (2009), River discharges of water and nutrients to the Mediterranean and Black Sea: Major drivers for ecosystem changes during past and future decades?, Progress in Oceanography, 80(3-4), 199-217. Luisetti, T., E. L. Jackson, and R. K. Turner (2013), Valuing the European coastal blue carbon storage benefit, Mar Pollut Bull, 71(1-2), 101-106. Lundsten, L., J. P. Barry, G. M. Cailliet, D. A. Clague, A. P. DeVogelaere, and J. B. Geller (2009), Benthic invertebrate communities on three seamounts off southern and central California, USA, Marine Ecology Progress Series, 374, 23-32. Ma, Z. L., Z. L. Xu, and J. Zhou (2009), Effect of global warming on the distribution of Lucifer intermedius and L. hanseni (Decapoda) in the Changjiang estuary, Progress in Natural Science, 19(10), 1389-1395. Mabus, R. (2010), Department of the Navy s Energy Program for Security and Independence, DIANE Publishing. Subject to Final Copyedit 94 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mackas, D. L., R. Goldblatt, and A. G. Lewis (1998), Interdecadal variation in developmental timing of Neocalanus plumchrus populations at Ocean Station P in the subarctic North Pacific, Canadian Journal of Fisheries and Aquatic Sciences, 55(8), 1878-1893. MacKenzie, B. R., and D. Schiedek (2007a), Daily ocean monitoring since the 1860s shows record warming of northern European seas, Global Change Biol, 13(7), 1335-1347. MacKenzie, B. R., and D. Schiedek (2007b), Long-term sea surface temperature baselines-time series, spatial covariation and implications for biological processes, J Marine Syst, 68(3-4), 405-420. MacKenzie, B. R., H. Gislason, C. Möllmann, and F. W. Köster (2007), Impact of 21st century climate change on the Baltic Sea fish community and fisheries, Global Change Biol, 13(7), 1348-1367. MacKenzie, B. R., H. E. M. Meier, M. Lindegren, S. Neuenfeldt, M. Eero, T. Blenckner, M. T. Tomczak, and S. Niiranen (2012), Impact of climate change on fish population dynamics in the Baltic Sea: A dynamical downscaling investigation, Ambio, 41, 626-636. MacNeil, M. A., N. A. Graham, J. E. Cinner, N. K. Dulvy, P. A. Loring, S. Jennings, N. V. Polunin, A. T. Fisk, and T. R. McClanahan (2010), Transitional states in marine fisheries: adapting to predicted global change, Philosophical Transactions of the Royal Society B: Biological Sciences, 365(1558), 3753-3763. Madsen, K. S., and N. K. Hjerslev (2009), Long-term temperature and salinity records from the Baltic Sea transition zone, Boreal Environ Res, 14(1), 125-131. Maier, C., J. Hegeman, M. G. Weinbauer, and J.-P. Gattuso (2009), Calcification of the cold-water coral Lophelia pertusa, under ambient and reduced pH, Biogeosciences, 6(8), 1671-1680. Maier, C., A. Schubert, M. M. B. Sanchez, M. G. Weinbauer, P. Watremez, and J.-P. Gattuso (2013), End of the century pCO2 levels do not impact calcification in Mediterranean cold-water corals, Plos One, 8(4), e62655. Makino, M., H. Matsuda, and Y. Sakurai (2009), Expanding fisheries co-management to ecosystem-based management: a case in the Shiretoko World Natural Heritage area, Japan, Mar Policy, 33(2), 207-214. Manhique, A. J., C. J. C. Reason, L. Rydberg, and N. Fauchereau (2011), ENSO and Indian Ocean sea surface temperatures and their relationships with tropical temperate troughs over Mozambique and the Southwest Indian Ocean, Int J Climatol, 31(1), 1-13. Mantua, N. J., and S. R. Hare (2002), The Pacific Decadal Oscillation, Journal of Oceanography, 58(1), 35-44. Mantua, N. J., S. R. Hare, Y. Zhang, J. M. Wallace, and R. C. Francis (1997), A Pacific interdecadal climate oscillation with impacts on salmon production, Bulletin of the American Meteorological Society, 78(6), 1069- 1079. Manzello, D. P. (2010), Coral growth with thermal stress and ocean acidification: lessons from the eastern tropical Pacific, Coral Reefs, 29(3), 749-758. Manzello, D. P., J. A. Kleypas, D. A. Budd, C. M. Eakin, P. W. Glynn, and C. Langdon (2008), Poorly cemented coral reefs of the eastern tropical Pacific: possible insights into reef development in a high-CO2 world, Proceedings of the National Academy of Sciences of the United States of America, 105(30), 10450-10455. Marba, N., and C. M. Duarte (2010), Mediterranean warming triggers seagrass (Posidonia oceanica) shoot mortality, Global Change Biol, 16(8), 2366-2375. Martin, J. H., and S. E. Fitzwater (1988), Iron deficiency limits phytoplankton growth in the north-east Pacific subarctic, Nature, 331, 341-343. Martin, S., and J.-P. Gatusso (2009), Response of Mediterranean coralline algae to ocean acidification and elevated temperature, Global Change Biol, 15(8), 2089-2100. Marullo, S., V. Artale, and R. Santoleri (2011), The SST multi-decadal variability in the Atlantic-Mediterranean region and its relation to AMO, Journal of Climate, 24(16), 4385-4401. Massuti, E., S. Monserrat, P. Oliver, J. Moranta, J. L. López-Jurado, M. Marcos, M. Hidalgo, B. Guijarro, A. Carbonell, and P. Pereda (2008), The influence of oceanographic scenarios on the population dynamics of demersal resources in the western Mediterranean: Hypothesis for hake and red shrimp off Balearic Islands, J Marine Syst, 71(3-4), 421-438. Matsuno, T., J. S. Lee, and S. Yanao (2009), The Kuroshio exchange with the South and East China Seas, Ocean Science, 5(3), 303-312. McAllister, C. D., T. R. Parsons, and J. D. H. Strickland (1960), Primary productivity and fertility at Station "P" in the North-East Pacific Ocean, ICES Journal of Marine Science, 25(3), 240-259. McCarthy, J. J., C. Garside, J. L. Nevins, and R. T. Barber (1996), New production along 140°W in the equatorial Pacific during and following the 1992 El Nino event, Deep-Sea Res Pt Ii, 43(4-6), 1065-1093. Subject to Final Copyedit 95 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McClain, C. R., S. R. Signorini, and J. R. Christian (2004), Subtropical gyre variability observed by ocean-color satellites, Deep-Sea Res Pt Ii, 51, 281-301. McClanahan, T., E. Weil, J. Cortés, A. H. Baird, and M. Ateweberhan (2009a), Consequences of coral bleaching for sessile reef organisms, in Coral Bleaching: Patterns, Processes, Causes and Consequences, edited by M. J. H. van Oppen and J. M. Lough, pp. 121-138, Springer-Verlag, Berlin & Heidelberg. McClanahan, T. R., E. Weil, and J. Maina (2009b), Strong relationship between coral bleaching and growth anomalies in massive Porites, Global Change Biol, 15(7), 1804-1816. McClanahan, T. R., M. Ateweberhan, J. Omukoto, and L. Pearson (2009c), Recent seawater temperature histories, status, and predictions for Madagascar's coral reefs, Marine Ecology Progress Series, 380, 117-128. McClanahan, T. R., M. Ateweberhan, N. A. J. Graham, S. K. Wilson, C. R. Sebastián, M. M. M. Guillaume, and J. H. Bruggemann (2007), Western Indian Ocean coral communities: bleaching responses and susceptibility to extinction, Marine Ecology Progress Series, 337, 1-13. McClanahan, T. R., et al. (2008), Conservation action in a changing climate, Conservation Letters, 1(2), 53-59. McClatchie, S., R. Goericke, R. Cosgrove, G. Auad, and R. Vetter (2010), Oxygen in the Southern California Bight: multidecadal trends and implications for demersal fisheries, Geophysical Research Letters, 37(19), L19602. McGowan, J. A., D. R. Cayan, and L. M. Dorman (1998), Climate-ocean variability and ecosystem response in the Northeast Pacific, Science, 281(5374), 210-216. McIlgorm, A., S. Hanna, G. Knapp, P. Le Floc H, F. Millerd, and M. Pan (2010), How will climate change alter fishery governance Insights from seven international case studies, Mar Policy, 34(1), 170-177. McKinnell, S. M., and M. J. Dagg (2010), Marine ecosystems of the North Pacific Ocean, 2003-2008, 393 pp. McKinnell, S. M., et al. (2010), Status and trends of the North Pacific Ocean, 2003-2008Rep., 1-55 pp. McLeod, E., R. Moffitt, A. Timmermann, R. Salm, L. Menviel, M. J. Palmer, E. R. Selig, K. S. Casey, and J. F. Bruno (2010), Warming seas in the Coral Triangle: Coral reef vulnerability and management implications, Coastal Management, 38(5), 518-539. McLeod, E., G. L. Chmura, S. Bouillon, R. Salm, M. Björk, C. M. Duarte, C. E. Lovelock, W. H. Schlesinger, and B. R. Silliman (2011), A blueprint for blue carbon: toward an improved understanding of the role of vegetated coastal habitats in sequestering CO2, Front Ecol Environ, 9(10), 552-560. McMichael, A. J., R. E. Woodruff, and S. Hales (2006), Climate change and human health: present and future risks, Lancet, 367(9513), 859-869. McNeil, B. I., and R. J. Matear (2008), Southern Ocean acidification: A tipping point at 450-ppm atmospheric CO2, Proceedings of the National Academy of Sciences of the United States of America, 105(48), 18860-18864. McPhaden, M. J., and D. Zhang (2002), Slowdown of the meridional overturning circulation in the upper Pacific Ocean, Nature, 415(6872), 603-608. Meehl, G. A., and J. M. Arblaster (2011), Decadal variability of Asian-Australian Monsoon-ENSO-TBO relationships, Journal of Climate, 24(18), 4925-4940. Meehl, G. A., A. X. Hu, and B. D. Santer (2009), The mid-1970s climate shift in the Pacific and the relative roles of rorced versus inherent decadal variability, Journal of Climate, 22(3), 780-792. Menard, F., F. Marsac, E. Bellier, and B. Cazelles (2007), Climatic oscillations and tuna catch rates in the Indian Ocean: a wavelet approach to time series analysis, Fisheries Oceanography, 16(1), 95-104. Mendelssohn, R., and F. Schwing (2002), Common and uncommon trends in SST and wind stress in the California and Peru Chile current systems, Progress in Oceanography, 53(2), 141-162. Menge, B. A., F. Chan, K. J. Nielsen, E. di Lorenzo, and J. Lubchenco (2009), Climatic variation alters supply-side ecology: impact of climate patterns on phytoplankton and mussel recruitment, Ecological Monographs, 79(3), 379-395. Merino, G., M. Barange, C. Mullon, and L. Rodwell (2010), Impacts of global environmental change and aquaculture expansion on marine ecosystems, Global Environmental Change-Human and Policy Dimensions, 20(4), 586-596. Merino, G., et al. (2012), Can marine fisheries and aquaculture meet fish demand from a growing human population in a changing climate?, Global Environmental Change, 22, 795-806. MerMex-Group, T. (2011), Marine ecosystems' responses to climatic and anthropogenic forcings in the Mediterranean, Progress in Oceanography, 91(2), 97-166. Mertz, O., K. Halsnaes, J. E. Olesen, and K. Rasmussen (2009), Adaptation to climate change in developing countries, Environmental Management, 43(5), 743-752. Subject to Final Copyedit 96 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Mestas-Nunez, A. M., and A. J. Miller (2006), Interdecadal variability and climate change in the eastern tropical Pacific: A review, Progress in Oceanography, 69(2-4), 267-284. Mieszkowska, N., S. J. Hawkins, M. T. Burrows, and M. A. Kendall (2007), Long-term changes in the geographic distribution and population structures of Osilinus lineatus (Gastropoda: Trochidae) in Britain and Ireland, J Mar Biol Assoc Uk, 87(2), 537-545. Millennium-Ecosystem-Assessment (2005), Ecosystems and human well-being: current state and trendsRep., Washington, Covelo and London. Miller, K. A., G. R. Munro, U. R. Sumaila, and W. W. L. Cheung (2013), Governing marine fisheries in a changing climate: a game-theoretic perspective, Canadian Journal of Agricultural Economics, 61, 309-334. Mills, D. J., L. Westlund, G. Graaf, Y. de Kura, R. Willman, and K. Kelleher (2011), Under-reported and undervalued: small-scale fisheries in the developing world, in Small-scale fisheries management: frameworks and approaches for the developing world, edited by R. S. Pomeroy and N. L. Andrew, pp. 1-15, CABI, Wallingford, UK. Minobe, S. (1997), A 50-70 year climatic oscillation over the North Pacific and North America, Geophysical Research Letters, 24(6), 683-686. Mohan, R., K. Verma, L. P. Mergulhao, D. K. Sinha, S. Shanvas, and M. V. S. Guptha (2006), Seasonal variation of pteropods from the Western Arabian Sea sediment trap, Geo-Marine Letters, 26(5), 265-273. Molinero, J. C., F. Ibanez, S. Souissi, M. Chifflet, and P. Nival (2005), Phenological changes in the Northwestern Mediterranean copepods Centropages typicus and Temora stylifera linked to climate forcing, Oecologia, 145(4), 640-649. Möllmann, C., G. Kornilovs, M. Fetter, and F. W. Köster (2005), Climate, zooplankton, and pelagic fish growth in the central Baltic Sea, ICES Journal of Marine Science, 62(7), 1270-1280. Möllmann, C., G. Kornilovs, M. Fetter, F. W. Köster, and H. H. Hinrichsen (2003), The marine copepod, Pseudocalanus elongatus, as a mediator between climate variability and fisheries in the Central Baltic Sea, Fisheries Oceanography, 12(4-5), 360-368. Möllmann, C., R. Diekman, B. Müller-Karulis, G. Kornilovs, M. Plikshs, and P. Axe (2009), Reorganization of a large marine ecosystem due to atmospheric and anthropogenic pressure: a discontinuous regime shift in the Central Baltic Sea, Global Change Biol, 15, 1377-1393. Monteiro, P. M. S., A. K. van der Plas, J. L. Mélice, and P. Florenchie (2008), Interannual hypoxia variability in a coastal upwelling system: Ocean-shelf exchange, climate and ecosystem-state implications, Deep Sea Research Part I: Oceanographic Research Papers, 55(4), 435-450. Morita, K., S. H. Morita, and M. A. Fukuwaka (2006), Population dynamics of Japanese pink salmon (Oncorhynchus gorbuscha): are recent increases explained by hatchery programs or climatic variations?, Canadian Journal of Fisheries and Aquatic Sciences, 63(5), 1189-1189. Moron, V. (2003), Long-term variability of the Mediterranean Sea surface temperature (1856-2000), C. R. Geoscience, 335(9), 721-727. Mountain, D. G., and J. Kane (2010), Major changes in the Georges Bank ecosystem, 1980s to the 1990s, Marine Ecology Progress Series, 398, 81-91. Movilla, J., E. Calvo, C. Pelejero, R. Coma, E. Serrano, P. Fernández-Vallejo, and M. Ribes (2012), Calcification reduction and recovery in native and non-native Mediterranean corals in response to ocean acidification, J Exp Mar Biol Ecol, 438, 144-153. Moy, A. D., W. R. Howard, S. G. Bray, and T. W. Trull (2009), Reduced calcification in modern Southern Ocean planktonic foraminifera, Nature Geoscience, 2(4), 276-280. Mueter, F. J., and M. A. Litzow (2008), Sea ice retreat alters the biogeography of the Bering Sea continental shelf, Ecol Appl, 18(2), 309-320. Muhling, B. A., S. K. Lee, J. T. Lamkin, and Y. Y. Liu (2011), Predicting the effects of climate change on bluefin tuna (Thunnus thynnus) spawning habitat in the Gulf of Mexico, ICES Journal of Marine Science, 68(6), 1051- 1062. Mumby, P. J., R. Vitolo, and D. B. Stephenson (2011), Temporal clustering of tropical cyclones and its ecosystem impacts, Proceedings of the National Academy of Sciences of the United States of America, 108(43), 17626- 17630. Mumby, P. J., J. R. M. Chisholm, A. J. Edwards, C. D. Clark, E. B. Roark, S. Andréfouët, and J. Jaubert (2001), Unprecedented bleaching-induced mortality in Porites spp. at Rangiroa Atoll, French Polynesia, Mar. Biol., 139(1), 183-189. Subject to Final Copyedit 97 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Munday, P. L., D. L. Dixson, J. M. Donelson, G. P. Jones, M. S. Pratchett, G. V. Devitsina, and K. B. Doving (2009), Ocean acidification impairs olfactory discrimination and homing ability of a marine fish, Proceedings of the National Academy of Sciences of the United States of America, 106(6), 1848-1852. Murray, J. W., H. W. Jannasch, S. Honjo, R. F. Anderson, W. S. Reeburgh, Z. Top, G. E. Friederich, L. A. Codispoti, and E. Izdar (1989), Unexpected changes in the oxic/anoxic interface in the Black Sea, Nature, 338(6214), 411-413. Myers, S. S., and J. A. Patz (2009), Emerging threats to human health from global environmental change, Annual Review of Environment and Resources, 34, 223-252. Nakamura, N., H. Kayanne, H. Iijima, T. R. McClanahan, S. K. Behera, and T. Yamagata (2011), Footprints of IOD and ENSO in the Kenyan coral record, Geophysical Research Letters, 38(24), L24708. Nakanowatari, T., K. I. Ohshima, and M. Wakatsuchi (2007), Warming and oxygen decrease of intermediate water in the northwestern North Pacific, originating from the Sea of Okhotsk, 1955-2004, Geophysical Research Letters, 34(4), L04602. Narayan, N., A. Paul, S. Mulitza, and M. Schulz (2010), Trends in coastal upwelling intensity during the late 20th century, Ocean Science, 6(3), 815-823. Naylor, R. L., et al. (2009), Feeding aquaculture in an era of finite resources, Proceedings of the National Academy of Sciences of the United States of America, 106, 15103-15110. Nellemann, C., and E. Corcoran (2009), Blue carbon: the role of healthy oceans in binding carbon: a rapid response assessment, UNEP/Earthprint. Nicol, S. J., et al. (2013), An ocean observation system for monitoring the affects of climate change on the ecology and sustainability of pelagic fisheries in the Pacific Ocean, Climatic Change, 119, 131-145. Ning, X., C. Lin, Q. Hao, C. Liu, F. Le, and J. Shi (2009), Long term changes in the ecosystem in the northern South China Sea during 1976-2004, Biogeosciences, 6(10), 2227-2243. Ning, X., C. Lin, J. Su, C. Liu, Q. Hao, and F. Le (2011), Long-term changes of dissolved oxygen, hypoxia, and the responses of the ecosystems in the East China Sea from 1975 to 1995, Journal of Oceanography, 67(1), 59-75. Norman-López, A., É. Plagányi, T. Skewes, E. Poloczanska, D. Dennis, M. T. Gibbs, and P. Bayliss (2013), Linking physiological, population and socio-economic assessments of climate-change impacts on fisheries, Fisheries Research. Nttestad, L., et al. (2013), Cruise report from the coordinated ecosystem survey (IESSMS) with M/V "Libas", M/V "Eros", M/V "Finnur Frí i" and R/V "Arni Fridriksson" in the Norwegian Sea and surrounding waters, 2 July - 9 August 2013, in Working Document to ICES Working Group on Widely distributed Stocks (WGWIDE), ICES Headquarters, Copenhagen, Denmark, 27 August - 2 September 2013, edited, p. 42. Nye, J. A., J. S. Link, J. A. Hare, and W. J. Overholtz (2009), Changing spatial distribution of fish stocks in relation to climate and population size on the Northeast United States continental shelf, Marine Ecology Progress Series, 393, 111-129. Obura, D., and S. Mangubhai (2011), Coral mortality associated with thermal fluctuations in the Phoenix Islands, 2002 2005, Coral Reefs, 1-13. Oguz, T., and V. Velikova (2010), Abrupt transition of the northwestern Black Sea shelf ecosystem from a eutrophic to an alternative pristine state, Marine Ecology Progress Series, 405, 231-242. Oguz, T., B. Fach, and B. Salihoglu (2008), Invasion dynamics of the alien ctenophore Mnemiopsis leidyi and its impact on anchovy collapse in the Black Sea, J Plankton Res, 30(12), 1385-1397. Oguz, T., T. Cokacar, P. Malanotte-Rizzoli, and H. W. Ducklow (2003), Climatic warming and accompanying changes in the ecological regime of the Black Sea during 1990s, Global Biogeochemical Cycles, 17(3), 1088. Ohmura, A. (2009), Observed decadal variations in surface solar radiation and their causes, Journal of Geophysical Research, 114, D00D05. Orr, J. C., et al. (2005), Anthropogenic ocean acidification over the twenty-first century and its impact on calcifying organisms, Nature, 437(7059), 681-686. Overland, J., S. Rodionov, S. Minobe, and N. Bond (2008), North Pacific regime shifts: Definitions, issues and recent transitions, Progress in Oceanography, 77(2-3), 92-102. Overland, J. E., J. Alheit, A. Bakun, J. W. Hurrell, D. L. Mackas, and A. J. Miller (2010), Climate controls on marine ecosystems and fish populations, J Marine Syst, 79(3-4), 305-315. Parmesan, C., C. Duarte, E. Poloczanska, A. J. Richardson, and M. C. Singer (2011), Commentary: Overstretching attribution, Nature Climate Change, 1(1), 2-4. Subject to Final Copyedit 98 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Parrish, R. H., F. B. Schwing, and R. Mendelssohn (2000), Mid-latitude wind stress: the energy source for climatic shifts in the North Pacific Ocean, Fisheries Oceanography, 9(3), 224-238. Pascual, M., X. Rodó, S. P. Ellner, R. Colwell, and M. J. Bouma (2000), Cholera dynamics and El Nino-Southern Oscillation, Science, 289(5485), 1766-1769. Patz, J. A., D. Campbell-Lendrum, T. Holloway, and J. A. Foley (2005), Impact of regional climate change on human health, Nature, 438(7066), 310-317. Pauly, D., and V. Christensen (1995), Primary production required to sustain global fisheries, Nature, 374, 255-257. Pecl, G., S. Frusher, C. Gardner, M. Haward, A. Hobday, S. Jennings, M. Nursey-Bray, A. Punt, H. Revill, and I. van Putten (2009), The east coast Tasmanian rock lobster fishery vulnerability to climate change impacts and adaptation response options. Report to the Department of Climate Change, AustraliaRep., 114 pp. Penaflor, E., W. Skirving, A. Strong, S. Heron, and L. David (2009), Sea-surface temperature and thermal stress in the Coral Triangle over the past two decades, Coral Reefs, 28(4), 841-850. Pendleton, L., D. C. Donato, B. C. Murray, S. Crooks, W. A. Jenkins, S. Sifleet, C. Craft, J. W. Fourqurean, J. B. Kauffman, and N. Marba (2012), Estimating global blue carbon emissions from conversion and degradation of vegetated coastal ecosystems, Plos One, 7(9), e43542. Pennington, J. T., K. L. Mahoney, V. S. Kuwahara, D. D. Kolber, R. Calienes, and F. P. Chavez (2006), Primary production in the eastern tropical Pacific: A review, Progress in Oceanography, 69(2-4), 285-317. Perry, A. L., P. J. Low, J. R. Ellis, and J. D. Reynolds (2005), Climate change and distribution shifts in marine fishes, Science, 308(5730), 1912-1915. Perry, R. I., P. Cury, K. Brander, S. Jennings, C. Möllmann, and B. Planque (2010), Sensitivity of marine systems to climate and fishing: Concepts, issues and management responses, J Marine Syst, 79, 427-435. Peterson, W. T., and F. B. Schwing (2003), A new climate regime in northeast Pacific ecosystems, Geophysical Research Letters, 30(17), 1896. Pickering, T., B. Ponia, C. Hair, P. C. Southgate, E. Poloczanska, L. Della Patrona, A. Teitelbaum, C. V. Mohan, M. Phillips, and J. D. Bell (2011), Vulnerability of aquaculture in the tropical Pacific to climate change, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. J. Hobday, pp. 647-732, Secretariat of the Pacific Community, Noumea, New Caledonia. Pierce, D. W., P. J. Gleckler, T. P. Barnett, B. D. Santer, and P. J. Durack (2012), The fingerprint of human-­ induced changes in the ocean's salinity and temperature fields, Geophysical Research Letters, 39(21), L21704. Pinnegar, J., T. Watt, and K. Kennedy (2012), Climate change risk assessment for the marine and fisheries sectorRep., 204 pp, DEFRA, London, UK. Pinsky, M. L., B. Worm, M. J. Fogarty, J. L. Sarmiento, and S. A. Levin (2013), Marine taxa track local climate velocities, Science, 341, 1239-1242. Pitcher, T. J., D. Kalikoski, K. Short, D. Varkey, and G. Pramod (2009), An evaluation of progress in implementing ecosystem-based management of fisheries in 33 countries, Mar Policy, 33(2), 223-232. Pitt, N. R., E. S. Poloczanska, and A. J. Hobday (2010), Climate-driven range changes in Tasmanian intertidal fauna, Marine and Freshwater Research, 61(9), 963-970. Plagányi, É. E., T. Skewes, M. Haddon, and N. Dowling (2013), Risk management tools for sustainable fisheries management under changing climate: a sea cucumber example, Climatic Change, 119, 181-197. Poloczanska, E., et al. (2013), Global imprint of climate change on marine life, Nature Climate Change, 3, 919-925. Poloczanska, E. S., S. J. Hawkins, A. J. Southward, and M. T. Burrows (2008), Modeling the response of populations of competing species to climate change, Ecology, 89(11), 3138-3149. Polovina, J. J., and P. A. Woodworth (2012), Declines in phytoplankton cell size in the subtropical oceans estimated from satellite remotely-sensed temperature and chlorophyll, 1998 2007, Deep Sea Research Part II: Topical Studies in Oceanography, 77-80(SI), 82-88. Polovina, J. J., E. A. Howell, and M. Abecassis (2008), Ocean's least productive waters are expanding, Geophysical Research Letters, 35(3), L03618. Polovina, J. J., J. P. Dunne, P. A. Woodworth, and E. A. Howell (2011), Projected expansion of the subtropical biome and contraction of the temperate and equatorial upwelling biomes in the North Pacific under global warming, ICES Journal of Marine Science, 68(6), 986-995. Pomeroy, R. S., and N. L. Andrew (2011), Small-scale fisheries management: frameworks and approaches for the developing world, CABI. Subject to Final Copyedit 99 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pörtner, H. O. (2002), Climate variations and the physiological basis of temperature dependent biogeography: systemic to molecular hierarchy of thermal tolerance in animals, Comparative Biochemistry and Physiology- Part A: Molecular & Integrative Physiology, 132(4), 739-761. Pörtner, H. O. (2010), Oxygen- and capacity-limitation of thermal tolerance: a matrix for integrating climate-related stressor effects in marine ecosystems, Journal of Experimental Biology, 213(6), 881-893. Pörtner, H. O., C. Bock, K. Deigweiher, and M. Lucassen (2007), Environmental change and the role of acid-base regulation in marine ectotherms, Comparative Biochemistry and Physiology-Part A: Molecular & Integrative Physiology, 148(33), S142-S142. Pörtner, H. O., et al. (2001), Climate induced temperature effects on growth performance, fecundity and recruitment in marine fish: developing a hypothesis for cause and effect relationships in Atlantic cod (Gadus morhua) and common eelpout (Zoarces viviparus), Continental Shelf Research, 21(18-19), 1975-1997. Power, S., T. Casey, C. Folland, A. Colman, and V. Mehta (1999), Inter-decadal modulation of the impact of ENSO on Australia, Clim Dynam, 15, 319-324. Power, S. B., and I. N. Smith (2007), Weakening of the Walker Circulation and apparent dominance of El Nino both reach record levels, but has ENSO really changed, Geophysical Research Letters, 34(18), L18702. Prakash, S., and R. Ramesh (2007), Is the Arabian Sea getting more productive?, Curr Sci India, 92(5), 667-670. Pratchett, M. S., P. Munday, and S. K. Wilson (2008), Effects of climate-induced coral bleaching on coral-reef fishes - ecological and economic consequences, Oceanography and Marine Biology: An Annual Review, 46, 251-296. Pratchett, M. S., A. S. Hoey, S. K. Wilson, V. Messmer, and N. A. J. Graham (2011a), Changes in biodiversity and functioning of reef fish assemblages following coral bleaching and coral loss, Diversity, 3(3), 424-452. Pratchett, M. S., et al. (2011b), Vulnerability of coastal fisheries in the tropical Pacific to climate change, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. J. Hobday, pp. 167-185, Secretariat of the Pacific Community, Noumea, New Caledonia. Przeslawski, R., I. Falkner, M. B. Ashcroft, and P. Hutchings (2012), Using rigorous selection criteria to investigate marine range shifts, Estuarine, Coastal and Shelf Science, 113, 205-212. Qi, Q., R. Cai, and Q. Zhang (2010), Low-frequency variability of the heat transport in source area of Kuroshio and its relation to SST in China seas, Journal of Oceanography in Taiwan Strait, 29(1), 106-113 (in Chinese, with English abstract). Qiu, Y. S., Z. J. Lin, and Y. Z. Wang (2010), Responses of fish production to fishing and climate variability in the northern South China Sea, Progress in Oceanography, 85(3-4), 197-212. Rabalais, N. N., R. E. Turner, R. J. Díaz, and D. Justiæ (2009), Global change and eutrophication of coastal waters, ICES Journal of Marine Science, 66(7), 1528-1537. Rabalais, N. N., R. J. Diaz, L. A. Levin, R. E. Turner, D. Gilbert, and J. Zhang (2010), Dynamics and distribution of natural and human-caused hypoxia, Biogeosciences, 7(2), 585-619. Racault, M. F., C. Le Quéré, E. Buitenhuis, S. Sathyendranath, and T. Platt (2012), Phytoplankton phenology in the global ocean, Ecological Indicators, 14(1), 152-163. Rahman, C. (2012), The implications of climate change for maritime security forces, in Climate Change and the Oceans; Gauging the Legal and Policy Currents in the Asia Pacific and Beyond, edited by R. Warner and C. Schofield, pp. 167-199, Edward Elgar Publishing, Cheltenham, UK; Northampton, MA, USA. Raitsos, D. E., I. Hoteit, P. K. Prihartato, T. Chronis, G. Triantafyllou, and Y. Abualnaja (2011), Abrupt warming of the Red Sea, Geophysical Research Letters, 38, L14601. Raitsos, D. E., G. Beaugrand, D. Georgopoulos, A. Zenetos, A. M. Pancucci-Papadopoulou, A. Theocharis, and E. Papathanassiou (2010), Global climate change amplifies the entry of tropical species into the Eastern Mediterranean Sea, Limnol Oceanogr, 55(4), 1478-1484. Randolph, M. F., C. Gaudin, S. M. Gourvenec, D. J. White, N. Boylan, and M. J. Cassidy (2010), Recent advances in offshore geotechnics for deep water oil and gas developments, Ocean Engineering, 38, 818 834. Raven, J., K. Caldeira, H. Elderfield, O. Hoegh-Guldberg, P. Liss, U. Riebesell, J. Shepherd, C. Turley, and A. Watson (2005), Ocean acidification due to increasing atmospheric carbon dioxideRep. 0854036172, 68 pp, The Royal Society, London, UK. Rayfuse, R., and R. Warner (2012), Climate change mitigation activities in the ocean: turning up the regulatory heat, in Climate Change and the Oceans: Gauging the Legal and Policy Currents in the Asia Pacific and Beyond, edited by R. Warner and C. Schofield, pp. 234-262, Edward Elgar Publishing, Cheltenham, UK; Northampton, MA, USA. Subject to Final Copyedit 100 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rayner, N. A., D. E. Parker, E. B. Horton, C. K. Folland, L. V. Alexander, D. P. Rowell, E. C. Kent, and A. Kaplan (2003), Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century, Journal of Geophysical Research, 108(D14), 4407. Reid, P. C., N. P. Holliday, and T. J. Smyth (2001), Pulses in the eastern margin current and warmer water off the north west European shelf linked to North Sea ecosystem changes, Marine Ecology Progress Series, 215, 283- 287. Reid, P. C., D. G. Johns, M. Edwards, M. Starr, M. Poulin, and P. Snoeijs (2007), A biological consequence of reducing Arctic ice cover: arrival of the Pacific diatom Neodenticula seminae in the North Atlantic for the first time in 800,000 years, Global Change Biol, 13(9), 1910-1921. Richter, I., S. K. Behera, Y. Masumoto, B. Taguchi, N. Komori, and T. Yamagata (2010), On the triggering of Benguela Ninos: Remote equatorial versus local influences, Geophysical Research Letters, 37(20), L20604. Ridgwell, A., and D. N. Schmidt (2010), Past constraints on the vulnerability of marine calcifiers to massive carbon dioxide release, Nature Geoscience, 3(3), 196-200. Riegl, B. (2002), Effects of the 1996 and 1998 positive sea-surface temperature anomalies on corals, coral diseases and fish in the Arabian Gulf (Dubai, UAE), Mar. Biol., 140(1), 29-40. Riegl, B. (2003), Climate change and coral reefs: different effects in two high-latitude areas (Arabian Gulf, South Africa), Coral Reefs, 22(4), 433-446. Riegl, B., and W. E. Piller (2003), Possible refugia for reefs in times of environmental stress, International Journal of Earth Sciences, 92(4), 520-531. Riegl, B. M., S. J. Purkis, A. S. Al-Cibahy, M. A. Abdel-Moati, and O. Hoegh-Guldberg (2011), Present limits to heat-adaptability in corals and population-level responses to climate extremes, Plos One, 6(9), e24802. Riegl, B. M., A. W. Bruckner, G. P. Rowlands, S. J. Purkis, and P. Renaud (2012), Red Sea coral reef trajectories over 2 decades suggest increasing community homogenization and decline in coral size, Plos One, 7(5), e38396. Rixen, M., et al. (2005), The Western Mediterranean Deep Water: A proxy for climate change, Geophysical Research Letters, 32(12), L12608. Robinson, J., P. Guillotreau, R. Jiménez-Toribio, F. Lantz, L. Nadzon, J. Dorizo, C. Gerry, and F. Marsac (2010), Impacts of climate variability on the tuna economy of Seychelles, Clim Res, 43(3), 149-162. Rockström, J., et al. (2009), A safe operating space for humanity, Nature, 461(7263), 472-475. Rodó, X., M. Pascual, G. Fuchs, and A. S. Faruque (2002), ENSO and cholera: a nonstationary link related to climate change?, Proceedings of the National Academy of Sciences of the United States of America, 99(20), 12901-12906. Rodolfo-Metalpa, R., C. Lombardi, S. Cocito, J. M. Hall-Spencer, and M. C. Gambi (2010), Effects of ocean acidification and high temperatures on the bryozoan Myriapora truncata at natural CO2 vents, Marine Ecology, 31(3), 447-456. Rodolfo-Metalpa, R., et al. (2011), Coral and mollusc resistance to ocean acidification adversely affected by warming, Nature Climate Change, 1(6), 308-312. Rogelj, J., M. Meinshausen, and R. Knutti (2012), Global warming under old and new scenarios using IPCC climate sensitivity range estimates, Nature Climate Change, 2, 248 253. Rouault, M., B. Pohl, and P. Penven (2010), Coastal oceanic climate change and variability from 1982 to 2009 around South Africa, African Journal of Marine Science, 32(2), 237-246. Roy, C., C. D. van der Lingen, J. C. Coetzee, and J. R. E. Lutjeharms (2007), Abrupt environmental shift associated with changes in the distribution of Cape anchovy Engraulis encrasicolus spawners in the southern Benguela, African Journal of Marine Science, 29(3), 309-319. Ruckelshaus, M., S. Doney, H. Galindo, J. Barry, F. Chan, J. Duffy, C. English, S. Gaines, J. Grebmeier, and A. Hollowed (2013), Securing ocean benefits for society in the face of climate change, Mar Policy, 40, 154-159. Rykaczewski, R. R., and J. P. Dunne (2010), Enhanced nutrient supply to the California Current Ecosystem with global warming and increased stratification in an earth system model, Geophysical Research Letters, 37, L21606. Sabatés, A., P. Martin, J. Lloret, and V. Raya (2006), Sea warming and fish distribution: the case of the small pelagic fish, Sardinella aurita, in the western Mediterranean, Global Change Biol, 12(11), 2209-2219. Sabine, C. L., et al. (2004), The oceanic sink for anthropogenic CO2, Science, 305(5682), 367-371. Subject to Final Copyedit 101 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Sakamoto, T. T., H. Hasumi, M. Ishii, S. Emori, T. Suzuki, T. Nishimura, and A. Sumi (2005), Responses of the Kuroshio and the Kuroshio Extension to global warming in a high-resolution climate model, Geophysical Research Letters, 32(14), L14617. Samuelsson, M. (1996), Interannual salinity variations in the Baltic Sea during the period 1954-1990, Continental Shelf Research, 16(11), 1463-1477. Santana-Casiano, J. M., M. González-Dávila, and I. R. Ucha (2009), Carbon dioxide fluxes in the Benguela upwelling system during winter and spring: A comparison between 2005 and 2006, Deep-Sea Res Pt Ii, 56(8- 10), 533-541. Sarmiento, J. L., et al. (2004), Response of ocean ecosystems to climate warming, Global Biogeochemical Cycles, 18(3), GB3003. Schluter, M. H., A. Merico, M. Reginatto, M. Boersma, K. H. Wiltshire, and W. Greve (2010), Phenological shifts of three interacting zooplankton groups in relation to climate change, Global Change Biol, 16(11), 3144-3153. Schofield, C., and A. Arsana (2012), Climate change and the limits of maritime jurisdiction, in Climate Change and the Oceans: Gauging the Legal and Policy Currents in the Asia Pacific and Beyond, edited by R. Warner and C. Schofield, pp. 127-152, Edward Elgar Publishing, Cheltenham, UK; Northampton, MA, USA. Schofield, O., S. M. Glenn, M. A. Moline, M. Oliver, A. Irwin, Y. Chao, and M. Arrott (2013), Ocean observatories and information: Building a global ocean observing network, in Earth System Monitoring, edited by J. Orcutt, pp. 319-336, Springer. Schyen, H., and S. Brathen (2011), The Northern Sea Route versus the Suez Canal: cases from bulk shipping, Journal of Transport Geography, 19, 977-983. Scott, D., M. C. Simpson, and R. Sim (2012a), The vulnerability of Caribbean coastal tourism to scenarios of climate change related sea level rise, Journal of Sustainable Tourism, 20(6), 883-898. Scott, D., C. M. Hall, and S. Gössling (2012b), Tourism and climate change: Impacts, adaptation and mitigation, 463 pp., Routledge, Abingdon, UK. Seibel, B. A., and P. J. Walsh (2001), Potential impacts of CO2 injection on deep-sea biota, Science, 294(5541), 319- 320. Serrano, E., R. Coma, M. Ribes, B. Weitzmann, García, M., and E. Ballesteros (2013), Rapid northward spread of a zooxanthellate coral enhanced by artificial structures and sea warming in the western Mediterranean, Plos One, 8(1), e52739. Shaffer, G., S. M. Olsen, and J. O. P. Pedersen (2009), Long-term ocean oxygen depletion in response to carbon dioxide emissions from fossil fuels, Nature Geoscience, 2(2), 105-109. Sherman, K., I. M. Belkin, K. D. Friedland, J. O'Reilly, and K. Hyde (2009), Accelerated warming and emergent trends in fisheries biomass yields of the world's large marine ecosystems, Ambio, 38(4), 215-224. Shoji, J., S. Toshito, K. Mizuno, Y. Kamimura, M. Hori, and K. Hirakawa (2011), Possible effects of global warming on fish recruitment: shifts in spawning season and latitudinal distribution can alter growth of fish early life stages through changes in daylength, ICES Journal of Marine Science, 68(6), 1165-1169. Signorini, S. R., and C. R. McClain (2012), Subtropical gyre variability as seen from satellites, Remote Sensing Letters, 3(6), 471-479. Signorini, S. R., R. G. Murtugudde, C. R. McClain, J. R. Christian, J. Picaut, and A. J. Busalacchi (1999), Biological and physical signatures in the tropical and subtropical Atlantic, Journal of Geophysical Research, 104(C8), 18367-18382. Silverman, J., B. Lazar, and J. Erez (2007), Effect of aragonite saturation, temperature, and nutrients on the community calcification rate of a coral reef, Journal of Geophysical Research, 112, C05004. Silverman, J., B. Lazar, L. Cao, K. Caldeira, and J. Erez (2009), Coral reefs may start dissolving when atmospheric CO2 doubles, Geophysical Research Letters, 36, L05606. Simpson, S. D., S. Jennings, M. P. Johnson, J. L. Blanchard, P. J. Schon, D. W. Sims, and M. J. Genner (2011), Continental shelf-wide response of a fish assemblage to rapid warming of the sea, Curr Biol, 21(18), 1565- 1570. Smith, C. R., F. C. De Leo, A. F. Bernardino, A. K. Sweetman, and P. M. Arbizu (2008), Abyssal food limitation, ecosystem structure and climate change, Trends Ecol Evol, 23(9), 518-528. Smith, J., K. L., H. A. Ruhl, B. J. Bett, D. S. Billett, R. S. Lampitt, and R. S. Kaufmann (2009), Climate, carbon cycling, and deep-ocean ecosystems, Proceedings of the National Academy of Sciences of the United States of America, 106(46), 19211-19218. Subject to Final Copyedit 102 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Smith, J., K.L., and R. S. Kaufmann (1999), Long-term discrepancy between food supply and demand in the deep eastern North Pacific, Science, 284(5417), 1174-1177. Smith, T., M. Brandt, J. Calnan, R. Nemeth, J. Blondeau, E. Kadison, M. Taylor, and P. Rothenberger (2013), Convergent mortality responses of Caribbean coral species to seawater warming, Ecosphere, 4(7), art87. Southward, A. (1980), The Western English Channel-an inconstant ecosystem?, Nature, 285, 361-366. Southward, A. J., S. J. Hawkins, and M. T. Burrows (1995), Seventy years' observations of changes in distribution and abundance of zooplankton and intertidal organisms in the western English Channel in relation to rising sea temperature, J Therm Biol, 20(1-2), 127-155. Southward, A. J., et al. (2005), Long-term oceanographic and ecological research in the western English Channel, Advances in Marine Biology, 47, 1-105. Stenevik, E. K., and S. Sundby (2007), Impacts of climate change on commercial fish stocks in Norwegian waters, Mar Policy, 31(1), 19-31. Stott, P. A., N. P. Gillett, G. C. Hegerl, D. J. Karoly, D. A. Stone, X. B. Zhang, and F. Zwiers (2010), Detection and attribution of climate change: a regional perspective, WIREs Climate Change, 1(2), 192-211. Stramma, L., G. C. Johnson, J. Sprintall, and V. Mohrholz (2008), Expanding oxygen-minimum zones in the tropical oceans, Science, 320(5876), 655-658. Stramma, L., S. Schmidtko, L. A. Levin, and G. C. Johnson (2010), Ocean oxygen minima expansions and their biological impacts, Deep-Sea Research Part I-Oceanographic Research Papers, 57(4), 587-595. Stramma, L., E. D. Prince, S. Schmidtko, J. Luo, J. P. Hoolihan, M. Visbeck, D. W. R. Wallace, P. Brandt, and A. Kortzinger (2011), Expansion of oxygen minimum zones may reduce available habitat for tropical pelagic fishes, Nature Clim. Change, advance online publication. Stramma, L., E. D. Prince, S. Schmidtko, J. Luo, J. P. Hoolihan, M. Visbeck, D. W. R. Wallace, P. Brandt, and A. Körtzinger (2012), Expansion of oxygen minimum zones may reduce available habitat for tropical pelagic fishes, Nature Climate Change, 2(1), 33-37. Strong, A. E., C. S. Barrientos, C. Duda, and J. Sapper (1997), Improved satellite technique for monitoring coral reef bleaching, paper presented at Proceedings of the 8th International Coral Reef Symposium, Smithsonian Tropical Research Institute, Panama. Strong, A. E., G. Liu, W. Skirving, and C. M. Eakin (2011), NOAA's Coral Reef Watch program from satellite observations, Annals of GIS, 17(2), 83-92. Sugisaki, H., M. Nonaka, S. Ishizaki, K. Hidaka, T. Kameda, Y. Hirota, Y. Oozeki, H. Kubota, and A. Takasuka (2010), Status and trends of the Kuroshio region, 2003-2008, in Marine Ecosystems of the North Pacific Ocean, 2003-2008. , edited by S. M. McKinnell and M. J. Dagg, pp. 330-359, PICES Special Publication. Sultan, N., P. Cochonat, J. P. Foucher, and J. Mienert (2004), Effect of gas hydrates melting on seafloor slope instability, Marine Geology, 213(1-4), 379-401. Sumaila, U. R., W. W. L. Cheung, V. W. Y. Lam, D. Pauly, and S. Herrick (2011), Climate change impacts on the biophysics and economics of world fisheries, Nature Climate Change, 1(9), 449-456. Sundby, S. (2000), Recruitment of Atlantic cod stocks in relation to temperature and advectlon of copepod populations, Sarsia, 85(4), 277-298. Sundby, S., and O. Nakken (2008), Spatial shifts in spawning habitats of Arcto-Norwegian cod related to multidecadal climate oscillations and climate change, ICES Journal of Marine Science, 65(6), 953-962. Sweatman, H., S. Delean, and C. Syms (2011), Assessing loss of coral cover on Australia s Great Barrier Reef over two decades, with implications for longer-term trends, Coral Reefs, 2, 521-531. Tallis, H., A. Guerry, and G. C. Daily (2013), Ecosystem services, in Encyclopedia of Biodiversity (Second Edition), edited by S. A. Levin, pp. 96-104, Academic Press, Waltham, UK. Tam, J., et al. (2008), Trophic modeling of the Northern Humboldt Current Ecosystem, Part I: Comparing trophic linkages under La Nina and El Nino conditions, Progress in Oceanography, 79(2-4), 352-365. Tang, Q. (2009), Changing states of the Yellow Sea Large Marine Ecosystem: Anthropogenic forcing and climate impacts, in Sustaining the World s Large Marine Ecosystems, edited by K. Sherman, M. C. Aquarone and S. Adams, pp. 77-88, IUCN, Gland, Switzerland. Tang, Q. S., X. S. Jin, J. Wang, Z. M. Zhuang, Y. Cui, and T. X. Meng (2003), Decadal-scale variations of ecosystem productivity and control mechanisms in the Bohai Sea, Fisheries Oceanography, 12(4-5), 223-233. Tang, X. H., F. Wang, Y. L. Chen, and M. K. Li (2009), Warming trend in northern East China Sea in recent four decades, Chin J Oceanol Limn, 27(2), 185-191. Subject to Final Copyedit 103 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Taning, A. V. (1949), On changes in the marine fauna of the North-Western Atlantic area, with special reference to Greenland, Rapports et Proces-Verbaux des Réunions/Conseil Permanent International pour l Exploration de la Mer, 125, 26-29. Taning, A. V. (1953), Long term changes in hydrography and fluctuations in fish stocksRep., 1-10 pp, International Commision for the Northwest Atlantic Fisheries, Halifax, NS, Canada. Tanzil, J. T. I., B. E. Brown, A. W. Tudhope, and R. P. Dunne (2009), Decline in skeletal growth of the coral Porites lutea from the Andaman Sea, South Thailand between 1984 and 2005, Coral Reefs, 28(2), 519-528. Tasker, M. L. (2008), The effect of climate change on the distribution and abundance of marine species in the OSPAR Maritime AreaRep., 45 pp, International Council for the Exploration of the Sea. Taylor, G. T., F. E. Muller-Karger, R. C. Thunell, M. I. Scranton, Y. Astor, R. Varela, L. T. Ghinaglia, L. Lorenzoni, K. A. Fanning, and S. Hameed (2012), Ecosystem responses in the southern Caribbean Sea to global climate change, Proceedings of the National Academy of Sciences of the United States of America, 109(47), 19315-19320. Taylor, M. H., et al. (2008), Trophic modeling of the Northern Humboldt Current Ecosystem, Part II: Elucidating ecosystem dynamics from 1995 to 2004 with a focus on the impact of ENSO, Progress in Oceanography, 79(2- 4), 366-378. ter Hofstede, R., J. G. Hiddink, and A. D. Rijnsdorp (2010), Regional warming changes fish species richness in the eastern North Atlantic Ocean, Marine Ecology Progress Series, 414, 1-9. Terray, L., M. E. Demory, M. Déqué, G. de Coetlogon, and E. Maisonnave (2004), Simulation of late-twenty-first- century changes in wintertime atmospheric circulation over Europe due to anthropogenic causes, Journal of Climate, 17(24), 4630-4635. Terray, L., L. Corre, S. Cravatte, T. Delcroix, G. Reverdin, and A. Ribes (2012), Near-surface salinity as nature's rain gauge to detect human influence on the tropical water cycle, Journal of Climate, 25(3), 958-977. Thackeray, S. J., T. H. Sparks, M. Frederiksen, S. Burthe, P. J. Bacon, J. R. Bell, M. S. Botham, T. M. Brereton, P. W. Bright, and L. Carvalho (2010), Trophic level asynchrony in rates of phenological change for marine, freshwater and terrestrial environments, Global Change Biol, 16(12), 3304-3313. Thornes, J., M. Rennie, H. Marsden, and L. Chapman (2012), Climate change risk assessment for the transport sectorRep., 166 pp, Department for the Environment, Food and Rural Affairs, UK, London, UK. Thresher, R. E., B. D. Tilbrook, S. Fallon, N. C. Wilson, and J. Adkins (2011), Effects of chronic low carbonate saturation levels on the distribution, growth and skeletal chemistry of deep-sea corals and other seamount megabenthos, Marine Ecology Progress Series, 442(2011), 87 99. Tian, Y. J., H. Kidokoro, T. Watanabe, Y. Igeta, H. Sakaji, and S. Ino (2012), Response of yellowtail, Seriola quinqueradiata, a key large predatory fish in the Japan Sea, to sea water temperature over the last century and potential effects of global warming, J Marine Syst, 91(1), 1-10. Tokinaga, H., and S. P. Xie (2011), Weakening of the equatorial Atlantic cold tongue over the past six decades, Nature Geoscience, 4(4), 222-226. Toresen, R., and O. J. Ostvedt (2000), Variation in abundance of Norwegian spring-­ spawning herring (Clupea harengus, Clupeidae) throughout the 20th century and the influence of climatic fluctuations, Fish Fish, 1(3), 231-256. Touratier, F., and C. Goyet (2011), Impact of the Eastern Mediterranean Transient on the distribution of anthropogenic CO2 and first estimate of acidification for the Mediterranean Sea, Deep-Sea Research Part I- Oceanographic Research Papers, 58(1), 1-15. Trenberth, K. E. (1990), Recent observed interdecadal climate changes in the Northern Hemisphere, Bulletin of the American Meteorological Society, 71(7), 988-993. Trenberth, K. E., and D. J. Shea (2006), Atlantic hurricanes and natural variability in 2005, Geophysical Research Letters, 33(12), L12704. Tsamenyi, M., and Q. Hanich (2012), Fisheries jurisdiction under the Law of the Sea Convention: rights and obligations in maritime zones under the sovereignty of Coastal States, The International Journal of Marine and Coastal Law, 27(4), 783-793. Tsikliras, A. (2008), Climate-related geographic shift and sudden population increase of a small pelagic fish (Sardinella aurita) in the eastern Mediterranean Sea, Marine Biology Research, 4(6), 477-481. Turley, C., K. Boot, J.-P. Gattuso, and L. Hansson (2011), The ocean acidification challenges facing science and society, Ocean acidification, 249-271. Subject to Final Copyedit 104 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Turner, R. E., and N. N. Rabalais (1994), Coastal eutrophication near the Mississippi river delta, Nature, 368(6472), 619-621. Turner, R. E., N. N. Rabalais, and D. Justic (2008), Gulf of Mexico hypoxia: alternate states and a legacy, Environmental Science & Technology, 42(7), 2323-2327. Turton, S., T. Dickson, W. Hadwen, B. Jorgensen, T. Pham, D. Simmons, P. Tremblay, and R. Wilson (2010), Developing an approach for tourism climate change assessment: evidence from four contrasting Australian case studies, Journal of Sustainable Tourism, 18(3), 429-447. Tyler, W. B., K. T. Briggs, D. B. Lewis, and R. G. Ford (1993), Seabird distribution and abundance in relation to oceanographic processes in the California Current System, in The Status, Ecology and Conservation of Marine Birds of the North Pacific. Canadian Wildlife Service Special Publication, edited by K. Vermeer, K. T. Briggs, K. H. Morgan and D. Siegel-Causey, pp. 48-60, Canadian Wildlife Service, Ottawa, Canada. Ukrainskii, V., and Y. I. Popov (2009), Climatic and hydrophysical conditions of the development of hypoxia in waters of the northwest shelf of the Black Sea, Physical Oceanography, 19(3), 140-150. UNCTAD (2009a), Review of Maritime Transport, 2009, in Report by the UNCTAD secretariat, edited, p. 221, United Nations, New York and Geneva. UNCTAD (2009b), Summary of Proceedings, in Multi-year expert meeting on transport and trade facilitation: Maritime transport and the climate change challenge, edited, p. 52, United Nations, Geneva, Switzerland. UNWTO, and UNEP (2008), Climate Change and Tourism Responding to Global ChallengesRep., 256 pp, World Tourism Organization and the United Nations Environment Programme, Madrid, Spain. Van Dover, C. L. (2012), Hydrothermal vent ecosystems and conservation, Oceanography, 25(1), 313-316. van Hooidonk, R., J. A. Maynard, and S. Planes (2013), Temporary refugia for coral reefs in a warming world, Nature Climate Change, 3, 508-511. van Vuuren, D. P., et al. (2011), The representative concentration pathways: an overview, Climatic Change, 109, 5- 31. Vantrepotte, V., and F. Mélin (2011), Inter-annual variations in the SeaWiFS global chlorophyll a concentration (1997-2007), Deep-Sea Research Part I-Oceanographic Research Papers, 58(4), 429-441. Vargas-Ángel, B., E. E. Looney, O. J. Vetter, and E. F. Coccagna (2011), Severe, widespread El Nino-associated coral bleaching in the US Phoenix Islands, B Mar Sci, 87(3), 623-638. Vargas-Yánez, M., F. Moya, M. C. Garcia-Martinez, E. Tel, P. Zunino, F. Plaza, J. Salat, J. Pascual, J. L. López- Jurado, and M. Serra (2010), Climate change in the Western Mediterranean Sea 1900-2008, J Marine Syst, 82(3), 171-176. Vecchi, G. A., B. J. Soden, A. T. Wittenerg, I. M. Held, A. Leetmaa, and M. J. Harrison (2006), Weakening of tropical Pacific atmospheric circulation due to anthropogenic forcing, Nature, 441(7089), 73-76. Veitch, J. (2007), The changing state of the Benguela Current Large Marine Ecosystem: Expert Workshop on climate change and variability and impacts thereof in the BCLME region, 15 16th May 2007. Climate Change workshop report. Available at http://www.dlist.org/document-library/the-changing-state-the-bclme-expert- workshop-climate-change-and-variability, edited, p. 55. Veron, J. E. N., L. M. Devantier, E. Turak, A. L. Green, S. Kininmonth, M. Stafford-Smith, and N. Peterson (2009), Delineating the coral triangle, Galaxea, Journal of Coral Reef Studies, 11(2), 91-100. Wang, C. Z., and S. F. Dong (2010), Is the basin-wide warming in the North Atlantic Ocean related to atmospheric carbon dioxide and global warming?, Geophysical Research Letters, 37(8), L08707. Wang, M., N. A. Bond, and J. E. Overland (2007a), Comparison of atmospheric forcing in four sub-arctic seas, Deep-Sea Res Pt Ii, 54(23-26), 2543-2559. Wang, M., J. E. Overland, V. Kattsov, J. E. Walsh, X. Zhang, and T. Pavlova (2007b), Intrinsic versus forced variation in coupled climate model simulations over the Arctic during the twentieth century, Journal of Climate, 20(6), 1093-1107. Wang, M. Y., J. E. Overland, D. B. Percival, and H. O. Mofjeld (2006a), Change in the arctic influence on Bering Sea climate during the twentieth century, Int J Climatol, 26(4), 531-539. Wang, X., J. R. Christian, R. Murtugudde, and A. J. Busalacchi (2006b), Spatial and temporal variability in new production in the equatorial Pacific during 1980-2003: Physical and biogeochemical controls, Deep-Sea Res Pt Ii, 53(5-7), 677-697. Warner, R. (2012), Oceans beyond boundaries: Environmental assessment frameworks, International Journal of Marine and Coastal Law, 27(2), 481-499. Subject to Final Copyedit 105 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Warner, R., and C. Schofield (2012), Climate Change and the Oceans: Gauging the Legal and Policy Currents in the Asia Pacific and Beyond, 274 pp., Edward Elgar Publishing, Cheltenham, UK; Northampton, MA, USA. Waycott, M., C. Collier, K. McMahon, P. Ralph, L. McKenzie, J. Udy, and A. Grech (2007), Vulnerability of seagrasses in the Great Barrier Reef to climate change, in Climate Change and the Great Barrier Reef: a vulnerability assessment, edited by J. E. Johnson and P. A. Marshall, pp. 193-236, Great Barrier Reef Marine Park Authority and Australian Greenhouse Office, Townsville, Queensland, Australia. Waycott, M., L. J. McKenzie, J. E. Mellors, J. E. Ellison, M. T. Sheaves, C. Collier, A.-M. Schwarz, A. Webb, J. E. Johnson, and C. E. Payri (2011), Vulnerability of mangroves, seagrasses and intertidal flats in the tropical Pacific to climate change, in Vulnerability of Tropical Pacific Fisheries and Aquaculture to Climate Change, edited by J. D. Bell, J. E. Johnson and A. J. Hobday, pp. 297-368, Secretariat of the Pacific Community, Noumea, New Caledonia. Weeks, S. J., B. Currie, and A. Bakun (2002), Satellite imaging: Massive emissions of toxic gas in the Atlantic, Nature, 415(6871), 493-494. West, J. M., S. H. Julius, P. Kareiva, C. Enquist, J. J. Lawler, B. Petersen, A. E. Johnson, and M. R. Shaw (2009), US natural resources and climate change: concepts and approaches for management adaptation, Environmental Management, 44(6), 1001-1021. Westmacott, S., H. Cesar, and L. Pet-Soede (2001), Socio-economic assessment of the impacts of the 1998 coral reef bleaching in the Indian Ocean: A summary, in Coral reef degradation in the Indian Ocean, edited by D. Souter, D. Obura and O. Lindén, pp. 143-159, CORDIO, Stockholm, Sweden. Wethey, D. S., and S. A. Woodin (2008), Ecological hindcasting of biogeographic responses to climate change in the European intertidal zone, Hydrobiologia, 606, 139-151. Whitney, F. A., H. J. Freeland, and M. Robert (2007), Persistently declining oxygen levels in the interior waters of the eastern subarctic Pacific, Progress in Oceanography, 75(2), 179-199. Wild, M. (2009), Global dimming and brightening: A review, Journal of Geophysical Research, 114(21), D00D16. Wilderbuer, T. K., A. B. Hollowed, W. J. Ingraham, P. D. Spencer, M. E. Conners, N. A. Bond, and G. E. Walters (2002), Flatfish recruitment response to decadal climatic variability and ocean conditions in the eastern Bering Sea, Progress in Oceanography, 55(1-2), 235-247. Wilkinson, C. (2004), Status of coral reefs of the World: 2004, 264 pp., Australian Institute of Marine Science, Townsville, Queensland, Australia. Wilkinson, C., and G. Hodgson (1999), Coral reefs and the 1997-1998 mass bleaching and mortality, Nature Resour, 35(2), 16-25. Wiltshire, K. H., and B. F. J. Manly (2004), The warming trend at Helgoland Roads, North Sea: phytoplankton response, Helgoland Mar Res, 58(4), 269-273. Wiltshire, K. H., A. M. Malzahn, K. Wirtz, W. Greve, S. Janisch, P. Mangelsdorf, B. F. J. Manly, and M. Boersma (2008), Resilience of North Sea phytoplankton spring bloom dynamics: An analysis of long-term data at Helgoland Roads, Limnol Oceanogr, 53(4), 1294-1302. Wong, A. P. S., N. L. Bindoff, and J. A. Church (2001), Freshwater and heat changes in the North and South Pacific Oceans between the 1960s and 1985-94, Journal of Climate, 14(7), 1613-1633. Wong, G. T. F., S. Y. Chao, Y. H. Li, and F. K. Shiah (2000), The Kuroshio edge exchange processes (KEEP) study - an introduction to hypotheses and highlights, Continental Shelf Research, 20(4-5), 335-347. Wood, K. R., and J. E. Overland (2010), Early 20th century Arctic warming in retrospect, Int J Climatol, 30(9), 1269-1279. Woodworth-Jefcoats, P. A., J. J. Polovina, J. P. Dunne, and J. L. Blanchard (2013), Ecosystem size structure response to 21st century climate projection: large fish abundance decreases in the central North Pacific and increases in the California Current, Global Change Biol, 19(3), 724-733. Worm, B., et al. (2009), Rebuilding global fisheries, Science, 325(5940), 578-585. WTTC (2012), The Economic Impact of Travel & Tourism 2012 Rep., 20 pp, World Travel & Tourism Council, London, UK. Wu, L. X., et al. (2012), Enhanced warming over the global subtropical western boundary currents, Nature Climate Change, 2(3), 161-166. Wunsch, C., and P. Heimbach (2008), How long to oceanic tracer and proxy equilibrium?, Quaternary Science Reviews, 27(7-8), 637-651. Yatsu, A., S. Chiba, Y. Yamanaka, S. Ito, Y. Shimizu, M. Kaeriyama, and Y. Watanabe (2013), Climate forcing and the Kuroshio/Oyashio ecosystem, ICES Journal of Marine Science, 70(5), 922-933. Subject to Final Copyedit 106 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Yatsu, A., K. Y. Aydin, J. R. King, G. A. McFarlane, S. Chiba, K. Tadokoro, M. Kaeriyama, and Y. Watanabe (2008), Elucidating dynamic responses of North Pacific fish populations to climatic forcing: Influence of life- history strategy, Progress in Oceanography, 77(2-3), 252-268. Ye, S., and X. Huang (2003), HABs in East China Sea: surveillance and monitoring, Marine Environmental Science/Haiyang Huanjing Kexue, 22(2), 10-14. Yoder, J. A., C. R. Mcclain, G. C. Feldman, and W. E. Esaias (1993), Annual cycles of phytoplankton chlorophyll concentrations in the global ocean - a satellite view, Global Biogeochemical Cycles, 7(1), 181-193. Yu, J. Y., and H. Y. Kao (2007), Decadal changes of ENSO persistence barrier in SST and ocean heat content indices: 1958 2001, Journal of Geophysical Research: Atmospheres (1984-2012), 112(D13106, doi:10.1029/2006JD007654). Yu, K. F., J. X. Zhao, Q. Shi, T. G. Chen, P. X. Wang, K. D. Collerson, and T. S. Liu (2006), U-series dating of dead Porites corals in the South China sea: Evidence for episodic coral mortality over the past two centuries, Quat Geochronol, 1(2), 129-141. Yunev, O. A., J. Carstensen, S. Moncheva, A. Khaliulin, G. AErtebjerg, and S. Nixon (2007), Nutrient and phytoplankton trends on the western Black Sea shelf in response to cultural eutrophication and climate changes, Estuar Coast Shelf S, 74(1-2), 63-76. Zeebe, R. E., and A. Ridgwell (2011), Past changes in ocean carbonate chemistry, in Ocean Acidification, edited by J.-P. Gattuso and L. Hansson, pp. 21-66, Oxford University Press, Oxford, UK. Zeeberg, J., A. Corten, P. Tjoe-Awie, J. Coca, and B. Hamady (2008), Climate modulates the effects of Sardinella aurita fisheries off Northwest Africa, Fisheries Research, 89(1), 65-75. Zervas, C. E. (2009), Sea Level Variations of the United States, 1854-2006, National Oceanic and Atmospheric Administration, US Dept. of Commerce, National Ocean Service, Center for Operational Oceanographic Products and Services, US Government, USA. Zhang, Q., Y. Hou, and T. Yan (2011), Inter-annual and inter-decadal variability of Kuroshio heat transport in the East China Sea, Int J Climatol, 32, 481-488. Zhang, X., J. A. Church, S. M. Platten, and D. Monselesan (2013), Projection of subtropical gyre circulation and associated sea level changes in the Pacific based on CMIP3 climate models, Clim Dynam, 1-14. Zhang, X., S. Shang, C. Zhang, Z. Lu, S. Shang, and X. Li (2005), Potential influence of sea surface temperature on the interannual fluctuation of the catch and the distribution of chub mackerel and round scad in the Minnan- Taiwan Bank fishing ground, Chinese Marine Science Bulletin, 24(4), 91-96. Zhao, M. X., K. F. Yu, Q. M. Zhang, Q. Shi, and G. J. Price (2012), Long-term decline of a fringing coral reef in the Northern South China Sea, Journal of Coastal Research, 28, 1088-1099. Subject to Final Copyedit 107 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 30-1: Regional changes in sea surface temperature (SST) over the period 1950 2009 using the Ocean regionalization specified in Figure 30-1a (for further detail of regions defined for analysis, see Figure SM30-1 and Table 30-2, column 1). A linear regression was fitted to the average of all 1×1 degree monthly SST data extracted from the HadISST1.1 data set [Rayner et al., 2003] for each sub-region over the period 1950 2009. All SST values less than -1.8oC, together with all SST pixels that were flagged as being sea ice, were reset to the freezing point of seawater (-1.8oC) to reflect the sea temperature under the ice. Separate analyses were also done to explore trends in the temperatures extracted from the coldest-ranked and the warmest-ranked month of each year (Table SM30-2). The table includes the slope of the regression (°C decade-1), the p-value for the slope being different from zero and the total change over 60 years (i.e., the slope of linear regression multiplied by 6 decades) for each category. The p-values that exceed 0.05 plus the associated slope and change values have a gray background, denoting the lower statistical confidence in the slope being different from zero (no slope). Note, changes with higher p-values may still describe informative trends although the level of confidence is lower that the slope is different from zero. Regression slope Total change over 60 years p value, slope different from zero °C/Decade °C/Decade °C/Decade Change Change Change °C/Decade °C/Decade °C/Decade Coolest All months Warmest over 60 over 60 over 60 Coolest All months Warmest Month Month years years (all years Month Month Sub-region (coolest months) (warmest month) month) 1. High Latitude Spring Bloom Systems (HLSBS) Indian Ocean 0.056 0.087 0.145 0.336 0.522 0.870 0.000 0.003 0.000 North Atlantic 0.054 0.073 0.116 0.324 0.438 0.696 0.001 0.15 0.000 South Atlantic 0.087 0.063 0.097 0.522 0.378 0.582 0.000 0.098 0.000 North Pacific (west) 0.052 0.071 0.013 0.312 0.426 0.078 0.52 0.403 0.462 North Pacific (east) 0.016 0.04 0.016 0.096 0.24 0.096 0.643 0.53 0.444 Total North Pacific 0.033 0.055 0.015 0.198 0.33 0.09 0.284 0.456 0.319 South Pacific (west) 0.043 0.017 0.044 0.258 0.102 0.264 0.016 0.652 0.147 South Pacific (east) 0.047 0.031 0.052 0.282 0.186 0.312 0.000 0.396 0.003 Total South Pacific 0.046 0.027 0.050 0.276 0.162 0.300 0.000 0.467 0.000 2. Equatorial Upwelling Systems (EUS) Atlantic Equatorial 0.101 0.090 0.079 0.606 0.540 0.474 0.000 0.000 0.000 Pacific Equatorial 0.079 0.071 0.065 0.474 0.426 0.39 0.096 0.001 0.071 3. Semi-Enclosed Seas (SES) Subject to Final Copyedit 108 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Arabian Gulf 0.027 0.099 0.042 0.162 0.594 0.252 0.577 0.305 0.282 Baltic Sea 0.352 0.165 0.06 2.112 0.99 0.36 0.000 0.155 0.299 Black Sea -0.004 0.053 0.139 -0.024 0.318 0.834 0.943 0.683 0.009 Mediterranean Sea 0.035 0.084 0.110 0.21 0.504 0.660 0.083 0.32 0.006 Red Sea 0.033 0.07 0.047 0.198 0.42 0.282 0.203 0.138 0.042 4. Coastal Boundary Systems (CBS) Western Atlantic 0.137 0.123 0.127 0.822 0.738 0.762 0.000 0.000 0.000 Caribbean/Gulf of Mexico 0.023 0.024 0.019 0.138 0.144 0.114 0.193 0.498 0.281 Western Indian Ocean 0.097 0.100 0.096 0.582 0.600 0.576 0.000 0.000 0.000 Eastern Indian Ocean 0.099 0.092 0.080 0.594 0.552 0.480 0.000 0.000 0.000 E Indian/SE Asia/W Pacific 0.144 0.134 0.107 0.864 0.804 0.642 0.000 0.000 0.000 5. Eastern Boundary Upwelling Ecosystems (EBUE) Benguela Current 0.062 0.032 0.002 0.372 0.192 0.012 0.012 0.437 0.958 California Current 0.117 0.122 0.076 0.702 0.732 0.456 0.026 0.011 0.125 Canary Current 0.054 0.089 0.106 0.324 0.534 0.636 0.166 0.014 0.000 Humboldt Current 0.051 0.059 0.104 0.306 0.354 0.624 0.285 0.205 0.013 6. Sub-Tropical Gyres Indian Ocean 0.141 0.112 0.103 0.846 0.672 0.618 0.000 0.000 0.000 North Atlantic 0.042 0.046 0.029 0.252 0.276 0.174 0.048 0.276 0.038 South Atlantic 0.079 0.083 0.098 0.474 0.498 0.588 0.000 0.017 0.000 North Pacific (west) 0.065 0.071 0.059 0.390 0.426 0.354 0.000 0.018 0.000 North Pacific (east) 0.008 0.042 0.051 0.048 0.252 0.306 0.617 0.133 0.014 Total North Pacific 0.034 0.055 0.051 0.204 0.33 0.306 0.001 0.053 0.000 South Pacific (west) 0.060 0.076 0.092 0.360 0.456 0.552 0.002 0.000 0.000 South Pacific (east) 0.055 0.056 0.088 0.330 0.336 0.528 0.000 0.058 0.000 Total South Pacific 0.056 0.060 0.089 0.336 0.360 0.534 0.000 0.027 0.000 Subject to Final Copyedit 109 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 7. Coral Reef Provinces (Figure 30-3) Caribbean/Gulf of Mexico 0.026 0.024 0.023 0.156 0.144 0.138 0.107 0.382 0.203 Coral Triangle & SE Asia 0.137 0.131 0.098 0.822 0.786 0.588 0.000 0.000 0.000 Eastern Indian Ocean 0.081 0.097 0.116 0.486 0.582 0.696 0.000 0.000 0.000 Western Indian Ocean 0.091 0.100 0.102 0.546 0.600 0.612 0.000 0.000 0.000 Eastern Pacific Ocean 0.079 0.094 0.101 0.474 0.564 0.606 0.106 0.000 0.023 Western Pacific Ocean 0.072 0.073 0.073 0.432 0.438 0.438 0.000 0.000 0.000 8. Basin Scale North Atlantic (combined) 0.045 0.061 0.090 0.270 0.366 0.540 0.002 0.198 0.000 South Atlantic (combined) 0.076 0.074 0.101 0.456 0.444 0.606 0.000 0.041 0.000 Atlantic Ocean Basin 0.060 0.068 0.091 0.360 0.408 0.546 0.000 0.000 0.000 Total North Pacific 0.030 0.052 0.046 0.180 0.312 0.276 0.000 0.248 0.006 Total South Pacific 0.055 0.048 0.075 0.330 0.288 0.450 0.000 0.115 0.000 Pacific Ocean Basin 0.043 0.052 0.046 0.258 0.312 0.276 0.000 0.000 0.006 Indian Ocean Basin 0.130 0.108 0.106 0.780 0.648 0.636 0.000 0.000 0.000 Subject to Final Copyedit 110 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 30-2: Examples of priority adaptation options and supporting policies to assist Pacific Island countries and territories to minimize the threats of climate change to the socio-economic benefits derived from pelagic and coastal fisheries and aquaculture, and to maximize the opportunities. These measures are classified as win-win (W-W) adaptations, which address other drivers of the sector in the short-term and climate change in the long-term, or lose- win (L-W) adaptations, where benefits exceed costs in the short-term but accrue under longer-term climate change (modified from [Bell et al., 2013b]). Adaptation options Supporting policies Economic development Full implementation of the vessel day scheme Strengthen national capacity to administer the VDS. (VDS) to control fishing effort by the Parties to the Adjust national tuna management plans and marketing Nauru Agreements (W-W). strategies to provide flexible arrangements to buy and Diversify sources of fish for canneries in the region sell tuna. and maintain trade agreements, e.g., an Economic Include implications of climate change in management Partnership Agreement with the European Union (W- objectives of the WCPFC. W). Apply national management measures to address Continued conservation and management measures climate change effects for subregional concentrations of for all species of tuna to maintain stocks at healthy tuna in archipelagic waters beyond the mandate of levels and make these valuable species more resilient WCPFC. to climate change (W-W). Require all industrial tuna vessels to provide Energy efficiency programmes to assist fleets to operational-level catch and effort data to improve the cope with oil price rises and minimise CO2 emissions, models for redistribution of tuna stocks during climate and reduce costs of fishing further afield as tuna change. distributionsshift east (W-W). Pan-Pacific tuna management through merger of the Western and Central Pacific Fisheries Commission (WCPFC) and Inter-American Tropical Tuna Commission to coordinate management measures across the tropical Pacific (L-W). Food security Manage catchment vegetation to reduce transfer of Strengthen governance for sustainable use of coastal sediments and nutrients to coasts to reduce damage to fish habitats by: (1) building national capacity to adjacent coastal coral reefs, mangroves and understand the threats of climate change; (2) seagrasses that support coastal fisheries (W-W). empowering communities to manage fish habitats; and Foster the care of coral reefs, mangroves and (3) changing agriculture, forestry and mining practices to seagrasses by preventing pollution, managing waste prevent sedimentation and pollution. and eliminating direct damage to these coastal fish Minimise barriers to landward migration of coastal habitats (W-W). habitats during development of strategies to assist other Provide for migration of fish habitats by prohibiting sectors to respond to climate change. construction adjacent to mangroves and seagrasses Apply primary fisheries management to stocks of and installing culverts beneath roads to help the coastal fish and shellfish to maintain their potential for plants colonise landward areas as sea level rises; (L- replenishment. W). Allocate the necessary quantities of tuna from total Sustain and diversify catches of demersal coastal national catches to food security to increase access to fish to maintain the replenishment potential of all fish for both urban and coastal populations. stocks (L-W). Dedicate a proportion of the revenue from fishing Increase access to tuna caught by industrial fleets licences to improve access to tuna for food security. through storing and selling tuna and by-catch landed Include anchored inshore FADs as part of national at major ports to provide inexpensive fish for rapidly infrastructure for food security. growing urban populations (W-W). Install fish aggregating devices (FADs) close to the coast to improve access to fish for rural communities Subject to Final Copyedit 111 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 as human populations increase and demersal fish decline (W-W). Develop coastal fisheries for small pelagic fish species, e.g. mackerel, anchovies, pilchards, sardines and scads (W-W?). Promote simple post-harvest methods, such as traditional smoking, salting and drying, to extend the shelf life of fish when abundant catches are landed (W-W). Livelihoods Relocate pearl farming operations to deeper water Provide incentives for aquaculture enterprises to assess and to sites closer to coral reefs and seagrass/algal risks to infrastructure so that farming operations and areas where water temperatures and aragonite facilities can be climate-proofed and relocated if saturation levels are likely to be more suitable for necessary. good growth and survival of pearl oysters, and Strengthen environmental impact assessments for formation of high-quality pearls (L-W). coastal aquaculture activities to include the additional Raise the walls and floor of shrimp ponds so that risks posed by climate change. they drain adequately as sea level rises (L-W). Develop partnerships with regional technical agencies Identify which shrimp ponds may need to be to provide support for development of sustainable rededicated to producing other commodities (L-W). aquaculture. a = The Parties to the Nauru Agreement (PNA) are Palau, Federated States of Micronesia, Papua New Guinea, Solomon Islands, Marshall Islands, Nauru, Kiribati, and Tuvalu. Subject to Final Copyedit 112 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 30-3: Key risks to ocean and coastal issues from climate change and the potential for risk reduction through mitigation and adaptation. Key risks are identified based on assessment of the literature and expert judgments made by authors of the various WGII AR5 chapters, with supporting evaluation of evidence and agreement in the referenced chapter sections. Each key risk is characterized as very low, low, medium, high, or very high. Risk levels are presented for the near-term era of committed climate change (here, for 2030 2040), in which projected levels of global mean temperature increase do not diverge substantially across emissions scenarios. Risk levels are also presented for the longer-term era of climate options (here, for 2080 2100), for global mean temperature increases of 2°C and 4°C above pre-industrial levels. For each timeframe, risk levels are estimated for the current state of adaptation and for a hypothetical highly adapted state. As the assessment considers potential impacts on different physical, biological, and human systems, risk levels should not necessarily be used to evaluate relative risk across key risks. Relevant climate variables are indicated by symbols. Subject to Final Copyedit 113 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 114 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 115 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 30-4: Ramifications, adaptation options and frameworks for decision-making for ocean regions. Symbols are as follows: T = sea temperature; UW = upwelling; OA = ocean acidification; NU = nutrient concentration; IC = ice cover; SS = storm strength; SLR = sea level rise ( = Increased; = decreased; italics = uncertain). Acronyms are: CBD (Convention on Biological Diversity); CTI (Coral Triangle Initiative); GEF (Global Environment Facility); IHO (International Hydrographic Organization); ILO (International Labor Organization); IOM (International Organization of Migration); ISPS (International Ship and Port Facility Security); MARPOL (International Convention for the Prevention of Pollution From Ships); PACC (Pacific Adaptation to Climate Change Project); PEMSEA (Partnerships in Environmental Management for the Seas of East Asia); RFMO (Regional Fisheries Management Organizations); SPREP (Secretariat of the Pacific Regional Environment Programme); UNCLOS (United Nations Convention on the Law of the Sea); UNHCR (United Nations High Commissioner for Refugees); UNSFSA (Straddling Fish Stocks Agreement); and WHO (World Health Organization). Primary Biophysical Key risks and Ramifications Adaptation options Policy frameworks and Key driver(s) change projected vulnerabilities initiatives (examples) References and Chapter sections T, OA Spatial and Reduced fisheries Reduced national Increased international cooperation LOSC, PEMSEA, CTI, [Tsamenyi and temporal variation production impacts income, increased over key fisheries. Improved RFMO agreements, Hanich, 2012] in primary important sources of unemployment, plus understanding of linkages between UNSFSA, [Bell et al., productivity income to some countries increase in poverty. ocean productivity, recruitment and 2011; Bell et (medium while others may see Potential increase in fisheries stock levels. Implementation al., 2013a] confidence at increase productivity (e.g., disputes over national of the regional vessel day scheme , 6.4.1, 6.5.3, global scales, Box as tuna stocks shift ownership of key fishery provide social and economic 30.6.2.1, CC-PP) eastwards in the resources (likely) incentives to fisheries and fishersfor 30.7.2, Box Pacific)(medium adaptation,. CC-PP, confidence). T, OA Ecosystem regime Reduced fisheries Decreased food and Strengthen coastal zone management PEMSEA, CTI, PACC, [Bell et al., shifts (e.g., coral to production as coastal employment security and to reduce contributing stressors (e.g., MARPOL, UNHCR, 2013a]; 5.4.3, algal reefs; habitats and ecosystems human migration away coastal pollution, over-harvesting and CBD, IOM, GEF, ILO 6.3.1-2, 12.4, structural shifts in such as coral reefs (medium from coastal zone (likely) physical damage to coastal resources). 30.5.2-4, phytoplankton confidence). Promote Blue Carbon* initiatives. 29.3.1, 29.3.3, communities, 30.5.6, 30.6.1, medium 30.6.2.1, Box confidence) CC-CR Tourist appeal of coastal Increased levels of As above, strengthen coastal zone CBD, PEMSEA, CTI, [Kenchington assets decreases as coastal poverty in some management and reduce additional PACC, UNHCR, and Warner, ecosystems change to less countries as tourist stressors on tourist sites; implement MARPOL 2012], 5.5.4.1, 'desirable' state reducing income decreases education programs and awareness 6.4.1 2, 10.6, income to some countries (likely). among visitors. Diversify tourism 30.6.2.2, (low confidence). activities. Subject to Final Copyedit 116 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Primary Biophysical Key risks and Ramifications Adaptation options Policy frameworks and Key driver(s) change projected vulnerabilities initiatives (examples) References and Chapter sections Increased risk of some Increased disease and Increased monitoring and education National policy [Llewellyn, diseases (e.g., ciguatera, mortality; decreases in surrounding key risks (e.g., strategies as well as and 2010] 6.4.2.3, harmful algal blooms) as coastal food resources ciguatera); develop alternate fisheries regional cooperation 10.6, 29.3.3.2, temperatures increase shift and fisheries income and income for periods when disease needed 29.5.3, 30.6.3, and ecosystems shift away (likely). incidence increases, develop or update from coral dominance (low health response plans. confidence) Increased poverty and Increased population Develop alternative industries and UNCLOS, PEMSEA, [Kaye, 2012; dislocation of coastal pressure on migration income for affected coastal people. CTI, ISPS, IMO, Bali Rahman, people (particularly in the destinations (e.g. large Strengthen coastal security both Process on Transnational 2012] tropics) as coastal regional cities), and nationally and across regions. Crime ASEAN Mutual 12.4-6,29.3.3, resources such as fisheries reduced freedom to Increase cooperation over criminal legal Assistance treaty 29.6.2, 30.6.5 degrade (medium navigate in some areas activities. and bilateral extradition confidence). (as criminal activity and MLA agreements increases; likely). T Migration of Reorganization of Social and economic Increased international cooperation UNCLOS, CBD, RFMO 7.4.2, 6.5, organisms and commercial fish stocks and disruption (very likely) and improve understanding of regime agreements, UNSFSA 30.5, 30.6.2.1, ecosystems to ecological regime shifts changes; early-detection monitoring Box CC-MB, higher latitudes (medium to high of physical and biological variables (high confidence). confidence). and regional seasonal forecasting; include related uncertainties into fisheries management; social and economic incentives for industry. Increased in abundance, Increased disease risk to Increase environmental monitoring; ; IMO, BWM, Anti 6.4.1.5, growing season and aquaculture and fisheries. technological advances to deal with Fouling Convention 7.3.2.4, distributional extent ofpests Income loss and pest and fouling organisms; increase 29.5.3-4, and fouling species increased operating and vigilance and control related to 30.6.2.1, Box (medium confidence). maintenance costs (very biosecurity. CC-MB likely) Threats to human health Increased disease and Reduce exposure through increased UNICEF, WHO, IHOs, [Myers and increase due to expansion mortality in some coastal monitoring and education, adoption or and national Patz, 2009]; of pathogens distribution to communities (likely) update of health response plans to governments. 6.4.3, 10.8.2, higher latitudes (low outbreaks 11.7, 29.3.3, confidence). 30.6.3, Box CC-MB Subject to Final Copyedit 117 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Primary Biophysical Key risks and Ramifications Adaptation options Policy frameworks and Key driver(s) change projected vulnerabilities initiatives (examples) References and Chapter sections T, NU, Increased Increased threats to Reduced supply of Provide early-detection monitoring CTI, PEMSEA, WHO, [Llewellyn, OA incidence of ecosystems, fisheries and marine fish and shellfish and improve predictive models, MARPOL 2010], 30.6.3, harmful algal human health (medium and greater incidence of provide education and adoption or 11.7, 6.4.2.3 blooms (HABs, confidence). disease among some update of health response plans. low confidence). coastal communities (likely). T Increased Increased freshwater, Increasing damage to Improve management of catchment CTI, PEMSEA, SPREP 3.4, 29.3.1, precipitation as a sediment and nutrients flow coastal reef systems with and coastal processes; expand riparian 30.5.4, 30.6.1 result of into coastal areas, increase ecological regime shifts vegetation along creeks and rivers; intensified in number and severity of in many cases (very improve agricultural retention of soils hydrological cycle flood events(medium to likely). and nutrients. in some coastal high confidence). areas (medium confidence) T Changing weather Increased risk of damage to Increased damage and Adjust infrastructure specifications, IMO [IPCC, 2012], patterns,, storm infrastructure such as that associated costs (likely) Developearly-warning systems and 10.4.4, 29.3, frequency , involved in shipping, and update emergency response plans to 30.6.2.3-4 medium oil and gas exploration and extreme events. confidence) extraction (medium to low confidence). SLR, Increased wave Exposure of coastal Increased costs to human Develop integrated coastal UNICEF, IHOs, and [Warner, SS exposure of coastal infrastructure and towns and settlements, management that consider SLR in national governments. 2012] areas and communities to damage numbers of displaced planning and decision-making; 5.5, 12.4.1, increased sea level and inundation, increase people and human increase understanding of the issues 29.5.1, (high confidence) coastal erosion (high migration (very likely). through education. 30.3.1.2, confidence) 30.6.5 Inundation of coastal Reduced food and water Assist communities to find UNICEF, IHOs, and [Warner, aquifers reduces water security leads to alternatives for food and water, or national governments. 2012], 5.4.3, supplies and decreased increased coastal assist in relocation of populations and 12.4.1, 29.3.2, coastal agricultural poverty, reduced food agriculture from vulnerable areas. 30.3.1.2 productivity (high security, and migration confidence). (very likely). Subject to Final Copyedit 118 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Primary Biophysical Key risks and Ramifications Adaptation options Policy frameworks and Key driver(s) change projected vulnerabilities initiatives (examples) References and Chapter sections SLR Risk of inundation UNCLOS defined limits of Lack of clarity increases Seek resolution of shifting national UNCLOS [IPCC, 2012; and coastal maritime jurisdiction will as do disputes over baselines issue (retreat and Schofield and erosion, especially contract as national maritime limits and redefinition, stabilization, or fixation Arsana, 2012; in low-lying baselines shift inland. maritime jurisdiction. of EEZ and other currently defined Warner and countries (high Potential uncertainty Some nations at risk of maritime jurisdiction limits. Schofield, confidence). increases in some areas major losses to their 2012] with respect to the territorial waters (very 5.5, 30.6.5 international boundaries to likely). maritime jurisdiction (high confidence). T, IC Loss of summer Access to northern coasts Potential for increased Seek early resolution of areas in UNCLOS Chapter 28 sea ice (high of Canada, USA and tension on different dispute currently and in the future. confidence) Russia increases security interpretations of access concerns (high confidence). rights and boundaries (likely to very likely). New resources become Tensions over maritime International agreements need to be UNCLOS Chapter 28 available as ice retreats, claims and ownership of sorted. increasing vulnerability of resources (likely). international borders in some cases (medium confidence). *Blue Carbon initiatives include conservation and restoration of mangroves, saltmarsh and seagrass beds as carbon sinks (30.6.1). Subject to Final Copyedit 119 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-1: (a) Separation of the world s non-polar oceans into seven major sub-regions (excluding the polar oceans, which are considered in Chapter 28). The chlorophyll-a signal measured by SeaWiFS and averaged over the period from Sep 4, 1997 to 30 Nov 2010 (NASA) is provides a proxy for differences in marine productivity (with the caveats provided in Box CC-PP). Ecosystem structure and functioning, as well as key oceanographic features provided the basis for separating the Ocean into the sub-regions shown. The map insert shows the distribution of Deep Sea (DS) habitat (>1000 m; Bathypelagic and Abyssopelagic habitats combined). Numbers refer to: 1 = High Latitude Spring Bloom Systems (HLSBS); 2 = Equatorial Upwelling Systems (EUS); 3 = Semi-Enclosed Seas (SES); 4 = Coastal Boundary Systems (CBS); 5 = Eastern Boundary Upwelling Ecosystems (EBUE); 6 = Sub- Tropical Gyres (STG); and 7 = DS (>1000 m). (b) Relationship between fish catch and area for each ocean sub- region is shown in (a). Red columns: average fish catch (as millions tons yr-1) for the period 1970 2006. Blue columns: area (millions km2). The four left-hand columns (sub-regions HLSBS-North, CBS, EBUE, and SES) cover 20 % of the world oceans area and deliver 80% of the world s fish catches. The values for the percent area of the Ocean, primary productivity, and fishery catch for the major sub-regions are listed in Table SM30-1. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 120 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-2: (a) Depth-averaged 0 700 m temperature trend for 1971 2010 (longitudinal versus latitude, colors and gray contours in °C per decade). (b) Zonally averaged temperature trends (latitude versus depth, colors and gray contours in °C per decade) for 1971 2010, with zonally averaged mean temperature over plotted (black contours in °C). (c) Globally-averaged temperature anomaly (Time versus depth, colors and grey contours in °C) relative to the 1971 2010 mean. (d) Globally-averaged temperature difference between the Ocean surface and 200 m depth (Black: annual values; red: five year running mean). Panels (a) (d) from WGI Figure 3.1. (e) (g) Observed and simulated variations in past and projected future annual average SST over three ocean basins (excluding regions within 300 km of the coast). The black line shows estimates from HadISST1.1 observational measurements. Shading denotes the 5 95 percentile range of climate model simulations driven with historical changes in anthropogenic and natural drivers (62 simulations), historical changes in natural drivers only (25), and the Representative Concentration Pathways: Dark Blue: RCP2.6; Light Blue: RCP4.5; Green: RCP6.0, and Red: RCP8.5). Data are anomalies from the 1986 2006 average of the HadISST1.1 data (for the HadISST1.1 time series) or of the corresponding historical all-forcing simulations. Further details are given in Box 21-2. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 121 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-3: Velocity at which sea surface temperature (SST) isotherms shifted (km decade-1) over the period 1960 2009 calculated using HaDISST1.1, with arrows indicating the direction and magnitude of shifts. Velocity of climate change is obtained by dividing the temperature trend in °C decade-1 by the local spatial gradient °C km-1. The direction of movement of SST is denoted by the direction of the spatial gradient and the sign of the temperature trend: towards locally cooler areas with a local warming trend or towards locally warmer areas where temperatures are cooling. Adapted from [Burrows et al., 2011]. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 122 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-4: Recent changes in thermal stress calculated using HadISST1.1 data. A monthly climatology was created by averaging the HadISST monthly SST values over the period 1985 2000 to create twelve averages, one for each month of the year. The Maximum Monthly Mean (MMM) climatology was created by selecting the hottest month for each pixel. Anomalies were then created by subtracting this value from each SST value, but only allowing values to be recorded if they were greater than zero [Donner et al., 2007]. Two measures of the change in thermal stress were calculated as a result: (a) The total thermal stress for the period 1981 2010, calculated by summing all monthly thermal anomalies for each grid cell. (b) The location of coral reef grid cells used in Table 30-1 and for comparison to regional heat stress here. Each dot is positioned over a 1×1 degree grid cell within which lies at least one carbonate coral reef. The latitude and longitude of each reef is derived from data provided by the World Resources Institute s Reefs at Risk report (http://www.wri.org). The six regions are as follows: Red Western Pacific Ocean; Yellow Eastern Pacific Ocean; Dark Blue Caribbean & Gulf of Mexico; Green Western Indian Ocean; Pink Eastern Indian Ocean; and Light Blue Coral Triangle & SE Asia. (c) Proportion of years with thermal stress, which is defined as any year that has a thermal anomaly, for the periods 1951 1980 and (d) 1981 2010. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 123 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-5. Map of the rate of change in sea surface height (geocentric sea level) for the period 1993 2012 derived from satellite altimetry. Also shown are relative sea level changes (gray lines) from selected tide gauge stations for the period 1950 2012. For comparison, an estimate of global mean sea level change is shown (red lines) with each tide gauge time series. The relatively large short-term oscillations in local sea level (gray lines) are due to the natural climate variability and ocean circulation. For example, the large regular deviations at Pago Pago are associated with the El Nino-Southern Oscillation. Figure originally presented in WGI FAQ 13.1, Figure 1). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 124 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-6: (a) The 1955 2005 climatological-mean sea surface salinity [Antonov et al., 2010] color contoured at 0.5 PSS78 intervals (black lines). (b) Annual mean evaporation-precipitation averaged over the period 1950 2000 (NCEP) color contoured at 0.5 m yr 1 intervals (black lines). (c) The 58-year (2008 minus 1950) sea surface salinity change derived from the linear trend (PSS78), with seasonal and ENSO signals removed [Durack and Wijffels, 2010] color contoured at 0.116 PSS78 intervals (black lines). (d) The 30-year (2003 2007 average centered at 2005, minus the 1960 1989 average centered at 1975) sea surface salinity difference (PSS78) color contoured at 0.06 PSS78 intervals (black lines). Contour intervals in (c) and (d) are chosen so that the trends can be easily compared, given the different time intervals in the two analyzes. White areas in (c) and (d) are marginal seas where the calculations are not carried out. Regions where the change is not significant at the 99% confidence level are stippled in gray. Figure originally presented as WGI Figure 3.4 in WGI. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 125 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-7: Projected ocean acidification from 11 CMIP5 Earth System models under RCP8.5 (other RCP scenarios have also been run with the CMIP5 models): (a) Time series of surface pH shown as the mean (solid line) and range of models (filled), given as area-weighted averages over the Arctic Ocean (green), the tropical oceans (red) and the Southern Ocean (blue). (b) Maps of the median model s change in surface pH from 1850 2100. Panel (a) also includes mean model results from RCP2.6 (dashed lines). Over most of the Ocean, gridded data products of carbonate system variables are used to correct each model for its present-day bias by subtracting the model-data difference at each grid cell following [Orr et al., 2005]. Where gridded data products are unavailable (Arctic Ocean, all marginal seas and the Ocean near Indonesia), the results are shown without bias correction. The bias correction reduces the range of model projections by up to a factor of 4, e.g., in panel (a) compare the large range of model projections for the Arctic (without bias correction) to the smaller range in the Southern Ocean (with bias correction). Figure originally presented in WGI Figure 6.28 in WGI. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 126 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-8: Projected aragonite saturation state from 11 CMIP5 Earth System models under RCP8.5 scenario: (a) time series of surface carbonate ion concentration shown as the mean (solid line) and range of models (filled), given as area weighted averages over the Arctic Ocean (green), the tropical oceans (red), and the Southern Ocean (blue); maps of the median model's surface A in (b) 2010, (d) 2050, and (f) 2100; and zonal mean sections (latitude versus depth) of A in 2100 over (c) the Atlantic Ocean and (e) the Pacific Ocean, while the ASH (Aragonite Saturation Horizon) is shown for 2010 (dotted line) and 2100 (solid line). Panel (a) also includes mean model results from RCP2.6 (dashed lines). As for Figure 30-7, gridded data products of carbonate system variables [Key et al., 2004] are used to correct each model for its present-day bias by subtracting the model-data difference at each grid cell following [Orr et al., 2005]. Where gridded data products are unavailable (Arctic Ocean, all marginal seas and the Ocean near Indonesia), results are shown without bias correction. Reprinted from Figure 6.29 in WGI. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 127 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-9: (a) Simulated changes in dissolved O2 (mean and model range as shading) relative to 1990s for RCP2.6, RCP4.5, RCP6.0, and RCP8.5. (b) Multi-model mean dissolved O2 (umol m 3) in the main thermocline (200 600 m depth average) for the 1990s, and changes in the 2090s relative to 1990s for RCP2.6 (c) and RCP8.5 (d). To indicate consistency in the sign of change, regions are stippled when at least 80% of models agree on the sign of the mean change. These diagnostics are detailed in [Cocco et al., 2013] in a previous model inter-comparison using the SRES-A2 scenario and have been applied to CMIP5 models here. Models used: CESM1-BGC, GFDL- ESM2G, GFDL-ESM2M, HadGEM2-ES, IPSL-CM5A-LR, IPSL-CM5A-MR, MPI-ESM-LR, MPI-ESM-MR, NorESM1. Figure originally presented in WGI Figure 6.30 in WGI. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 128 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Subject to Final Copyedit 129 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-10: Annual maximum proportions of reef pixels with Degree Heating Months [Donner et al., 2007]; DHM) 1 (used for projecting coral bleaching; [Strong et al., 1997; Strong et al., 2011]) and DHM 5 (associated with bleaching across 100% of affected areas with significant mortality, [Eakin et al., 2010] for the period 1870 2009 for each of the six coral regions (Figure 30-4d) using the HadISST1.1 data set. The black line on each graph is the maximum annual area value for each decade over the period 1870 2009. This value is continued through 2010 2099 using CMIP5 data and splits into the four Representative Concentration Pathways (RCP2.6, 4.5, 6.0, and 8.5). DHM were produced for each of the four RCPs using the ensembles of CMIP models. From these global maps of DHM, the annual percentage of grid cells with DHM 1 and DHM 5 were calculated for each coral region. These data were then grouped into decades from which the maximum annual proportions were derived. The plotted lines for 2010 2099 are the average of these maximum proportion values for each RCP. Monthly SST anomalies were derived using a 1985 2000 maximum monthly mean (MMM) climatology derived in the calculations for Figure 30- 4. This was done separately for HadISST1.1, the CMIP5 models, and each of the four RCPs, at each grid cell for every region. DHMs were then derived by adding up the monthly anomalies using a 4-month rolling sum. Figure SM30-3 presents past and future sea temperatures for the six major coral reef provinces under historic, un-forced, RCP4.5 and RCP8.5 scenarios.] Subject to Final Copyedit 130 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-11: Expert assessment of degree of confidence in detection and attribution of physical and chemical changes (a) and ecological changes (b) across sub-regions, as designated in Figure 30-1a, and processes in the Ocean (based on evidence explored throughout Chapter 30 and elsewhere in AR5). Further explanation of this figure is given in 18.3.3 4 and 18.6. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 131 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 30-12: (a) Examples of projected impacts and vulnerabilities associated with climate change in Ocean sub-regions. (b) Examples of risks to fisheries from observed and projected impacts across Ocean sub-regions. Letters indicate level of confidence: (vL): Very low, (L): Low, (M): Medium, (H): High and (vH): Very high. Details of sub-regions are given in Table 30-1a and 30.1.1. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 132 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure CR-1: A and B: the same coral community before and after a bleaching event in February 2002 at 5 m depth, Halfway Island, Great Barrier Reef. Coral cover at the time of bleaching was 95% bleached almost all of it severely bleached, resulting in mortality of 20.9% (Elvidge et al., 2004). Mortality was comparatively low due in part because these coral communities were able to shuffle their symbiont to more thermo-tolerant types (Berkelmans and van Oppen, 2006; Jones et al., 2008). C and D: three CO2 seeps in Milne Bay Province, Papua New Guinea show that prolonged exposure to high CO2 is related to fundamental changes in the ecology of coral reefs (Fabricius et al., 2011), including reduced coral diversity (-39%), severely reduced structural complexity (-67%), lower density of young corals (-66%) and fewer crustose coralline algae (-85%). At high CO2 sites (panel D; median pHT ~7.8), reefs are dominated by massive corals while corals with high morphological complexity are underrepresented compared with control sites (D; median pH ~8.0). Reef development ceases at pHT values below 7.7. pHT: pH on the total scale. E: temporal trend in coral cover for the whole Great Barrier Reef over the period 1985 2012 (N, number of reefs, mean +/- 2 standard errors; De'ath et al., 2012). F: composite bars indicate the estimated mean coral mortality for each year, and the sub-bars indicate the relative mortality due to crown-of-thorns starfish, cyclones, and bleaching for the whole Great Barrier Reef (De'ath et al., 2012). Photo credit: R. Berkelmans (A and B) and K. Fabricius (C and D). Subject to Final Copyedit 133 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure MB-1: 1735 observed responses to climate change from 208 single- and multi-species studies. Changes attributed to climate change (blue), inconsistent with climate change (red) and are equivocal (yellow). Each circle represents the centre of a study area. Where points fall on land, it is because they are centroids of distribution that surround an island or peninsula. Pie charts show the proportions within regions bounded by red squares and in the Mediterranean; numbers indicate the total (consistent, opposite or equivocal) observations within each region. Note: 57% of the studies included were published since AR4 (from Poloczanska et al., 2013). [Illustration to be redrawn to conform to IPCC publication specifications.] Figure MB-2. Rates of change in distribution (km decade-1) for marine taxonomic groups, measured at the leading edges (red) and trailing edges (brown). Average distribution shifts calculated using all data, regardless of range location, are in black. Distribution rates have been square-root transformed; standard errors may be asymmetric as a result. Positive distribution changes are consistent with warming (into previously cooler waters, generally poleward). Means +/- standard error are shown, along with number of observations (from Poloczanska et al., 2013). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 134 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure MB-3. A. Rate of climate change for the Ocean (sea surface temperature (SST) °C); B. corresponding climate velocities for the Ocean and median velocity from land (adapted from Burrows et al., 2011); and C. observed rates of displacement of marine taxonomic groups over several decades until 2010. The thin dotted red arrows give an example of interpretation. Rates of climate change of 0.008 °C yr-1 correspond to ca. 2.4 km yr-1median climate velocity in the Ocean. When compared to observed rates of displacement, many marine taxonomic groups have been able to track these velocities, except phyto- and zooplankton where rates of displacement greatly exceed climate velocity. All values are calculated for ocean surface with the exclusion of polar seas (Figure 30-1a). (A) Observed rates of climate change for Ocean SST (Black dotted line) are derived from HadISST1.1 data set, all other rates are calculated based on the average of the CMIP5 climate model ensembles (Table S30-3) for the historical period and for the future based on the four RCP emissions scenarios. Data were smoothed using a 20-year sliding window. (B) Median climate velocity calculated from HadISST1.1 dataset over 1960 2010 using the methods of Burrows et al., 2011. The three axes represent estimated median climate velocities are representative of areas of slow velocities such as Pacific subtropical gyre (STG) system (Purple line), the global Ocean surface (excluding polar seas, Blue line), and areas of high velocities such as the Coral Triangle and North Sea (Orange line). Figure 30-3 shows climate velocities over the ocean surface calculated over 1960 2010. The Red line corresponds to the median rate over global land surface calculated using historical surface temperatures from the CMIP5 model ensemble (Table S30-3). (C) Rates of displacement for marine taxonomic groups estimated by Poloczanska et al. 2013 using published studies (Figure MB-2 Black data set). Note the displacement rates for phytoplankton exceed the axis, so values are given. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 135 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure OA-1: A: Overview of the chemical, biological, socio-economic impacts of ocean acidification and of policy options (adapted from Turley and Gattuso, 2012). B: Multi-model simulated time series of global mean ocean surface pH (on the total scale) from CMIP5 climate model simulations from 1850 to 2100. Projections are shown for emission scenarios RCP2.6 (blue) and RCP8.5 (red) for the multi-model mean (solid lines) and range across the distribution of individual model simulations (shading). Black (grey shading) is the modelled historical evolution using historical reconstructed forcings. The models that are included are those from CMIP5 that simulate the global carbon cycle while being driven by prescribed atmospheric CO2 concentrations. The number of CMIP5 models to calculate the multi-model mean is indicated for each time period/scenario (WGI AR5 Figure 6.28). C: Effect of near future acidification (seawater pH reduction of 0.5 unit or less) on major response variables estimated using weighted random effects meta-analyses, with the exception of survival which is not weighted (Kroeker et al., 2013). The log- transformed response ratio (LnRR) is the ratio of the mean effect in the acidification treatment to the mean effect in a control group. It indicates which process is most uniformly affected by ocean acidification but large variability exists between species. Significance is determined when the 95% bootstrapped confidence interval does not cross zero. The number of experiments used in the analyses is shown in parentheses. * denotes a statistically significant effect. Subject to Final Copyedit 136 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure PP-1: A) Environmental factors controlling Net Primary Production (NPP). NPP is mainly controlled by three basic processes: 1) Light conditions in the surface ocean, i.e. the photic zone where photosynthesis occurs, 2) upward flux of nutrients and micronutrients from underlying waters into the photic zone, 3) Regeneration of nutrients and micronutrients via the breakdown and recycling of organic material before it sinks out of the photic zone. All three processes are influenced by physical, chemical and biological processes and vary across regional ecosystems. In addition, water temperature strongly influences the upper rate of photosynthesis for cells that are resource-replete. Predictions of alteration of primary productivity under climate change depend on correct parameterizations and simulations of each of these variables and processes for each region. B) Annual composite map of global areal NPP rates (derived from MODIS Aqua satellite climatology from 2003-2012; NPP was calculated with the Carbon-based Production Model (CbPM, Westberry et al., 2008)). Overlaid is a grid of (thin black lines) that represent 51 distinct global ocean biogeographical provinces (after Longhurst, 1998 and based on Boyd and Doney, 2002). The characteristics and boundaries of each province are primarily set by the underlying regional ocean physics and chemistry. Figure courtesy of Toby Westberry (OSU) and Ivan Lima (WHOI), satellite data courtesy of NASA Ocean Biology Processing Group. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 137 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 30 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure UP-1: Upper panel: Schematic hypothetic mechanism of increasing coastal wind-driven upwelling at eastern boundary systems, where differential warming rates between land and ocean results in increased land-ocean pressure gradients (1) that produce stronger alongshore winds (2) and offshore movement of surface water through Ekman transport (3), and increased upwelling of deep cold nutrient rich waters to replace it (4). Lower panel: potential consequences of climate change in upwelling systems. Increasing stratification and uncertainty in wind stress trends result in uncertain trends in upwelling. Increasing upwelling may result in higher input of nutrients to the euphotic zone, and increased primary production, which in turn may enhance pelagic fisheries, but also decreased coastal fisheries due to an augmented exposure of coastal fauna to hypoxic, low pH waters. Decreased upwelling may result in lower primary production in these systems with direct impacts on pelagic fisheries productivity. Subject to Final Copyedit 138 28 October 2013