FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter 11. Human Health: Impacts, Adaptation, and Co-Benefits Coordinating Lead Authors Kirk R. Smith (USA), Alistair Woodward (New Zealand) Lead Authors Diarmid Campbell-Lendrum (WHO), Dave Chadee (Trinidad and Tobago), Yasushi Honda (Japan), Qiyong Liu (China), Jane Olwoch (South Africa), Boris Revich (Russian Federation), Rainer Sauerborn (Sweden) Contributing Authors Clara Aranda (Mexico), Helen Berry (Australia), Colin Butler (Australia), Zoë Chafe (USA), Lara Cushing (USA), Kristie Ebi (USA), Tord Kjellstrom (New Zealand), Sari Kovats (UK), Graeme Lindsay (New Zealand), Erin Lipp (USA), Tony McMichael (Australia), Virginia Murray (UK), Osman Sankoh (Sierra Leone), Marie O’Neill (USA), Seth B. Shonkoff (USA), Joan Sutherland (Trinidad and Tobago), Shelby Yamamoto (Germany) Review Editors Ulisses Confalonieri (Brazil), Andrew Haines (UK) Volunteer Chapter Scientists Zoë Chafe (USA), Joachim Rocklov (Sweden) Contents Executive Summary 11.1. Introduction 11.1.1. Present State of Global Health 11.1.2. Developments since AR4 11.1.3. Non-Climate Health Effects of Climate-Altering Pollutants (CAPs) 11.2. How Climate Change Affects Health 11.3. Vulnerability to Disease and Injury due to Climate Variability and Climate Change 11.3.1. Geographic Causes of Vulnerability 11.3.2. Current Health Status 11.3.3. Age and Gender 11.3.4. Socioeconomic Status 11.3.5. Public Health and Other Infrastructure 11.3.6. Projections for Vulnerability 11.4. Direct Impacts of Climate and Weather on Health 11.4.1. Heat and Cold – Related Impacts 11.4.1.1. Mechanisms 11.4.1.2. Near-Term Future 11.4.2. Floods and Storms 11.4.2.1. Mechanisms 11.4.2.2. Near-Term Future 11.4.3. Ultraviolet Radiation 11.5. Ecosystem-Mediated Impacts of Climate Change on Health Outcomes 11.5.1. Vector-Borne and Other Infectious Diseases 11.5.1.1. Malaria 11.5.1.2. Dengue Fever Subject to Final Copyedit 1 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 11.5.1.3. Tick-Borne Diseases 11.5.1.4. Other Vector-Borne Diseases 11.5.1.5. Near-Term Future 11.5.2. Food- and Water-Borne Infections 11.5.2.1. Vibrios 11.5.2.2. Other Parasites, Bacteria, and Viruses 11.5.2.3. Near-Term Future 11.5.3. Air Quality 11.5.3.1. Long-Term Outdoor Ozone Exposures 11.5.3.2. Acute Air Pollution Episodes 11.5.3.3. Aeroallergens 11.5.3.4. Near-Term Future 11.6. Health Impacts Heavily Mediated through Human Institutions 11.6.1. Nutrition 11.6.1.1. Mechanisms 11.6.1.2. Near-Term Future 11.6.2. Occupational Health 11.6.2.1. Heat Strain and Heat Stroke 11.6.2.2. Heat Exhaustion and Work Capacity Loss 11.6.2.3. Other Occupational Health Concerns 11.6.2.4. Near-Term Future 11.6.3. Mental Health 11.6.4. Violence and Conflict 11.7. Adaptation to Protect Health 11.7.1. Improving Basic Public Health and Health Care Services 11.7.2. Health Adaptation Policies and Measures 11.7.3. Early Warning Systems 11.7.4. Role of Other Sectors in Health Adaptation 11.8. Adaptation Limits under High Levels of Warming 11.8.1. Physiological Limits to Human Heat Tolerance 11.8.2. Limits to Food Production and Human Nutrition 11.8.3. Thermal Tolerance of Disease Vectors 11.8.4. Displacement and Migration under Extreme Warming 11.8.5. Reliance on Infrastructure 11.9. Co-Benefits 11.9.1. Reduction of Co-Pollutants 11.9.1.1. Outdoor Sources 11.9.1.2. Household Sources 11.9.1.3. Primary Co-Pollutants 11.9.1.4. Secondary Co-Pollutants 11.9.1.5. Case Studies of Co-Benefits of Air Pollution Reductions 11.9.2. Access to Reproductive Health Services 11.9.2.1. Birth and Pregnancy Intervals 11.9.2.2. Maternal Age at Birth 11.10. Key Uncertainties and Knowledge Gaps References Subject to Final Copyedit 2 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Chapter Boxes 11-1. Weather, Climate, and Health – a Long-Term Observational Study in African and Asian Populations 11-2. Case Study: An Intervention to Control Dengue Fever 11-3. Health and Economic Impacts of Climate-Altering Pollutants (CAPs) Other than CO2 Frequently Asked Questions 11.1: How does climate change affect human health? 11.2: Will climate change have benefits for health? 11.3: Who is most affected by climate change? 11.4: What is the most important adaptation strategy to reduce the health impacts of climate change? 11.5: What are health “co-benefits” of climate change mitigation measures? Executive Summary The health of human populations is sensitive to shifts in weather patterns and other aspects of climate change (very high confidence). These effects occur directly, due to changes in temperature and precipitation and occurrence of heat waves, floods, droughts, and fires. Indirectly, health may be damaged by ecological disruptions brought on by climate change (crop failures, shifting patterns of disease vectors), or social responses to climate change (such as displacement of populations following prolonged drought). Variability in temperatures is a risk factor in its own right, over and above the influence of average temperatures on heat-related deaths. [11.4] Biological and social adaptation is more difficult in a highly variable climate than one that is more stable. [11.7] Until mid-century climate change will act mainly by exacerbating health problems that already exist [very high confidence]. New conditions may emerge under climate change [low confidence], and existing diseases (e.g. food- borne infections) may extend their range into areas that are presently unaffected [high confidence]. But the largest risks will apply in populations that are currently most affected by climate-related diseases. Thus, for example, it is expected that health losses due to climate change-induced under-nutrition will occur mainly in areas that are already food-insecure. [11.3] In recent decades, climate change has contributed to levels of ill-health (likely) though the present world-wide burden of ill-health from climate change is relatively small compared with other stressors on health and is not well quantified. Rising temperatures have increased the risk of heat-related death and illness (likely). [11.4] Local changes in temperature and rainfall have altered distribution of some water-borne illnesses and disease vectors, and reduced food production for some vulnerable populations [medium confidence]. [11.5, 11.6] If climate change continues as projected across the RCP scenarios until mid-century, the major increases of ill-health compared to no climate change will occur through: • Greater risk of injury, disease, and death due to more intense heat waves and fires [very high confidence] [11.4] • Increased risk of under-nutrition resulting from diminished food production in poor regions [high confidence] [11.6] • Consequences for health of lost work capacity and reduced labor productivity in vulnerable populations [high confidence] [11.6] • Increased risks of food- and water-borne diseases [very high confidence] and vector-borne diseases [medium confidence] [11.5] • Modest improvements in cold-related mortality and morbidity in some areas due to fewer cold extremes [low confidence], geographical shifts in food production, and reduced capacity of disease-carrying vectors due to exceedance of thermal thresholds [medium confidence]. These positive effects will be out-weighed, world-wide, by the magnitude and severity of the negative effects of climate change [high confidence]. [11.4, 11.5, 11.6] Impacts on health will be reduced, but not eliminated, in populations that benefit from rapid social and economic development [high confidence], particularly among the poorest and least healthy groups [very high confidence]. Subject to Final Copyedit 3 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 [11.4, 11.6, 11.7] Climate change is an impediment to continued health improvements in many parts of the world. If economic growth does not benefit the poor, the health effects of climate change will be exacerbated. In addition to their implications for climate change, essentially all the important Climate Altering Pollutants (CAPs) other than CO2 have near-term health implications [very high confidence]. In 2010, more than 7% of the global burden of disease was due to inhalation of these air pollutants [high confidence]. [Box 11-4] Some parts of the world already exceed the international standard for safe work activity during the hottest months of the year. The capacity of the human body to thermoregulate may be exceeded on a regular basis, particularly during manual labour, in parts of the world during this century. In the highest Representative Concentration Pathway, RCP8.5, by 2100 some of the world’s land area will be experiencing 4-7 degree higher temperatures due to anthropogenic climate change [WG1, Figure SPM.7]. If this occurs, the combination of high temperatures and high humidity will compromise normal human activities, including growing food or working outdoors, raising doubt about the habitability of some areas, for parts of the year [high confidence]. [11.8] The most effective adaptation measures for health in the near-term are programs that implement basic public health measures such as provision of clean water and sanitation, secure essential health care including vaccination and child health services, increase capacity for disaster preparedness and response, and alleviate poverty [very high confidence]. [11.7] In addition, there has been progress since AR4 in targeted and climate-specific measures to protect health, including enhanced surveillance and early warning systems. [11.7] There are opportunities to achieve co-benefits from actions that reduce emissions of CAPs and at the same time improve health. Among others, these include: • Reducing local emissions of health-damaging and climate-altering air pollutants from energy systems, through improved energy efficiency, and a shift to cleaner energy sources [very high confidence] [11.9] • Providing access to reproductive health services (including modern family planning) to improve child and maternal health through birth spacing and reduce population growth, energy use, and consequent CAP emissions over time [medium confidence] [11.9] • Shifting consumption away from animal products, especially from ruminant sources, in high-meat- consumption societies toward less CAP-intensive healthy diets [medium confidence] [11.9] • Designing transport systems that promote active transport and reduce use of motorized vehicles, leading to lower emissions of CAPs and better health through improved air quality and greater physical activity [high confidence] [11.9]. There are important research gaps regarding the health consequences of climate change and co-benefits actions, particularly in low-income countries. There are now opportunities to use existing longitudinal data on population health to investigate how climate change affects the most vulnerable populations. Another gap concerns the scientific evaluation of the health implications of adaptation measures at community and national levels. A further challenge is to improve understanding of the extent to which taking health co-benefits into account can offset the costs of GHG mitigation strategies. 11.1. Introduction This chapter examines what is known about the effects of climate change on human health and, briefly, the more direct impacts of Climate-Altering Pollutants (CAPs, see glossary) on health. We review diseases and other aspects of poor health that are sensitive to weather and climate. We examine the factors that influence the susceptibility of populations and individuals to ill-health due to variations in weather and climate, and describe steps that may be taken to reduce the impacts of climate change on human health. The chapter also includes a section on health “co- benefits.” Co-benefits are positive effects on human health that arise from interventions to reduce emissions of CAPs or vice versa. This is a scientific assessment based on best available evidence according to the judgment of the authors. We searched the English-language literature up to August 2013, focusing primarily on publications since 2007. We drew Subject to Final Copyedit 4 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 primarily (but not exclusively) on peer-reviewed journals. Literature was identified using a published protocol (Hosking and Campbell-Lendrum, 2012) and other approaches, including extensive consultation with technical experts in the field. We examined recent substantial reviews (for instance (Bassil and Cole, 2010; Gosling et al., 2009; Hajat et al., 2010; Huang et al., 2011; McMichael, 2013b; Stanke et al., 2013)), to check for any omissions of important work. In selecting citations for the chapter, we gave priority to publications that were recent (since AR4), comprehensive, added significant new findings to the literature, included areas or population groups that have not previously been well-described or were judged to be particularly policy-relevant in other respects. We begin with an outline of measures of human health, the major driving forces that act on health world-wide, recent trends in health status, and health projections for the remainder of this century. 11.1.1. Present State of Global Health The Fourth Assessment Report pointed to dramatic improvement in life expectancy in most parts of the world in the 20th Century, and this trend has continued through the first decade of the 21st century (Wang, 2012). Rapid progress in a few countries (especially China) has dominated global averages, but most countries have benefited from substantial reductions in mortality. There remain sizable and avoidable inequalities in life expectancy within- and between-nations in terms of education, income and ethnicity (Beaglehole and Bonita, 2008) and in some countries, official statistics are so patchy in quality and coverage that it is difficult to draw firm conclusions about health trends (Byass, 2010). Years lived with disability have tended to increase in most countries (Salomon et al., 2012). If economic development continues as forecast, it is expected that mortality rates will continue to fall in most countries; WHO estimates the global burden of disease (measured in Disability Adjusted Life Years per capita) will decrease by 30% by 2030, compared with 2004 (World Health Organization, 2008a; World Health Organization, 2008b). The underlying causes of global poor health are expected to change substantially, with much greater prominence of chronic diseases and injury, nevertheless the major infectious diseases of adults and children will remain important in some regions, particularly Sub-Saharan Africa and South Asia (Hughes et al., 2011). 11.1.2. Developments since AR4 The relevant literature has grown considerably since publication of AR4. For instance, the annual number of MEDLINE citations on climate change and health doubled between 2007 and 2009 (Hosking and Campbell- Lendrum, 2012). In addition, there have been many reviews, reports and international assessments that do not appear in listings such as MEDLINE but include important information nevertheless, for instance, the World Development Report 2010 (World Bank, 2010) and the 2011 UN Habitat report on cities and climate change (United Nations Human Settlements Programme, 2011). Since AR4, there have been improvements in the methods applied to investigate climate change and health. These include more sophisticated modeling of possible future impacts (for example, work linking climate change, food security, and health outcomes) (Nelson et al., 2010) and new methods to model the effects of heat on work capacity and labor productivity (Kjellstrom et al., 2009b). Other developments include coupling of high-quality, longitudinal mortality data sets with down-scaled meteorological data, in low- income settings (for instance, through the INDEPTH Network) (see Box 11-1). _____ START BOX 11-1 HERE _____ Box 11-1. Weather, Climate, and Health – a Long-Term Observational Study in African and Asian Populations Given the dearth of scientific evidence of the relationship between weather/climate and health in low- and middle- income countries, we report on a project that spans sub-Saharan Africa and Asia. The INDEPTH Network currently includes 43 surveillance sites in 20 countries. Using standardized health and demographic surveillance systems, member sites have collected up to 45 years information on births, migration and deaths. Currently, there are about 3.2 million people under surveillance (Sankoh and Byass, 2012). Subject to Final Copyedit 5 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 To study relationships between weather and health, the authors obtained daily meteorological data for 12 INDEPTH populations between 2000 and 2009, and projected future climate changes to 2100 under the A1B, A3, and B1 scenarios (Hondula et al., 2012). The authors concluded the health of all the populations would be challenged by the new climatic conditions, especially later in the century. In another study from the Network, Diboulo et al. (2012) examined the relation between weather and all-cause mortality data in Burkina Faso. Relations between daily temperature and mortality were similar to those reported in many high-income settings, and susceptibility to heat varied by age and gender. _____ END BOX 11-1 HERE _____ Since AR4, studies of the ways in which policies to reduce GHG emissions may affect health, or vice versa, leading to so-called “co-benefits” in the case of positive outcomes for either climate or health, have multiplied (Haines et al., 2009). Much has been written on links between climate, socioeconomic conditions and health, for example related to occupational heat exposure (Kjellstrom et al., 2009b) and malaria (e.g., Béguin et al., 2011; Gething et al., 2010) There is also growing appreciation of the social upheaval and damage to population health that may arise from the interaction of large-scale food insecurity, population dislocation, and conflict (see Chapter 12). 11.1.3. Non-Climate Health Effects of Climate-Altering Pollutants (CAPs) CAPs affect health in other ways than through climate change, just as CO2 creates non-climate effects such as ocean acidification. The effects of rising CO2 levels on calcifying marine species are well documented and the risks for coral reefs are now more closely defined than they were at the time of the AR4 (see Chapter 30). There are potential implications for human health, such as under-nutrition in coastal populations that depend on local fish stocks, but, so far, links between health and ocean acidification have not been closely studied (Kite-Powell et al., 2008). CAPs such as black carbon and tropospheric ozone have substantial, direct, negative effects on human health (Wang et al., 2013). (See 11.5.3 and Box 11-3.) Although CO2 is not considered a health-damaging air pollutant at levels experienced outside particular occupational and health-care settings, one study has reported a reduction in mental performance at 1000 ppm and above, within the range that all of humanity would experience in some extreme climate scenarios by 2100 (Satish et al., 2012). 11.2. How Climate Change Affects Health There are three basic pathways by which climate change affects health (Figure 11-1), and these provide the organization for the chapter: • Direct impacts, which relate primarily to changes in the frequency of extreme weather including heat, drought, and heavy rain [11.4] • Effects mediated through natural systems, for example, disease vectors, water-borne diseases, and air pollution [11.5] • Effects heavily mediated by human systems, for example, occupational impacts, undernutrition, and mental stress [11.6]. [INSERT FIGURE 11-1 HERE Figure 11-1: Conceptual diagram showing three primary exposure pathways by which climate change affects health: directly through weather variables such as heat and storms; indirectly through natural systems such as disease vectors; and pathways heavily mediated through human systems such as undernutrition. The yellow box indicates the moderating influences of local environmental conditions on how climate change exposure pathways are manifest in a particular population. The orange box indicates that the extent to which the three categories of exposure translate to actual health burden is moderated by such factors as background public health and socioeconomic conditions, and adaptation measures. The green arrows at the bottom indicate that there may be feedback Subject to Final Copyedit 6 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 mechanisms, positive or negative, between societal infrastructure, public health, and adaptation measures and climate change itself. As discussed later in the chapter, for example, some measures to improve health also reduce emissions of climate-altering pollutants, thus reducing the extent and/or pace of climate change as well as improving local health. Credit: E. Garcia, UC Berkeley.] The negative effects of climate change on health may be reduced by improved health services, better disaster management, and poverty alleviation, although the cost and effort may be considerable [11.7]. The consequences of large magnitude climate change beyond 2050, however, would be much more difficult to deal with [11.8]. Although there are exceptions, to a first approximation climate change acts to exacerbate existing patterns of ill-health, by acting on the underlying vulnerabilities that lead to ill-health even without climate change. Thus, before pursuing the three pathways in Figure 11-1, we summarize what is known about vulnerability to climate-induced illness and injury. 11.3. Vulnerability to Disease and Injury due to Climate Variability and Climate Change In the IPCC assessments, vulnerability is defined as the propensity or predisposition to be adversely affected (see Chapter 19 and Glossary). In this section, we consider causes of vulnerability to ill-health associated with climate change and climate variability, including individual and population characteristics and factors in the physical environment. We have outlined the causes of vulnerability separately, but in practice causes combine, often in a complex and place-specific manner. There are some factors (such as education, income, health status and responsiveness of government) that act as generic causes of vulnerability. For example, the quality of governance – how decisions are made and put into practice – affects a community’s response to threats of all kinds (Bowen et al., 2012). (See Chapter 12.) The background climate-related disease rate of a population is often the best single indicator of vulnerability to climate change - doubling of risk of disease in a low disease population has much less absolute impact than doubling of the disease when the background rate is high. (Note that here, and elsewhere in the chapter, we treat “risk” in the epidemiological sense: the probability that an event will occur.) But the precise causes of vulnerability, and therefore the most relevant adaptation capacities, vary greatly from one setting to another. For example, severe drought in Australia has been linked to psychological distress – but only for those residing in rural and remote areas (Berry et al., 2010). The link between high ambient temperatures and increased incidence of salmonella food poisoning has been demonstrated in many places (e.g. (Zhang et al., 2010) ), but the lag varies from one country to another, suggesting that the mechanisms differ. Deficiencies in food storage may be the critical link in some places, food handling problems may be most important elsewhere (Kovats et al., 2004). The 2010 World Development Report concluded that all developing regions are vulnerable to economic and social damage resulting from climate change – but for different reasons (World Bank, 2010). The critical factors for Sub- Saharan Africa, for example, are the current climate stresses (in particular, droughts and floods) that may be amplified in parts of the region under climate change, sparse infrastructure and high dependence on natural resources (see Chapter 22). Asia and the Pacific, on the other hand, are distinguished by the very large number of people living in low-lying areas prone to flooding (see Chapters 24 and 29). 11.3.1. Geographic Causes of Vulnerability Location has an important influence on the potential for health losses caused by climate change (Samson et al., 2011). Those working outdoors in countries where temperatures in the hottest time of the year are already at the limits of thermal tolerance for part of the year will be more severely affected by further warming than workers in cooler countries (Kjellstrom et al., 2013). The inhabitants of low-lying coral atolls are very sensitive to flooding, contamination of fresh water reservoirs due to sea level rise, and salination of soil, all of which may have important effects on health (Nunn, 2009). Rural populations that rely on subsistence farming in low rainfall areas are at high risk of under-nutrition and water-related diseases if drought occurs, although this vulnerability may be modified strongly by local factors, such as access to markets and irrigation facilities (Acosta-Michlik et al., 2008). Living in Subject to Final Copyedit 7 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 rural and remote areas may confer increased risk of ill-health because of limited access to services and generally higher levels of social and economic disadvantage (Smith, 2008). Populations that are close to the present limits of transmission of vector-borne diseases are most vulnerable to changes in the range of transmission due to rising temperatures and altered patterns of rainfall, especially when disease control systems are weak (Lozano-Fuentes et al., 2012; Zhou et al., 2008). In cities, those who live on urban heat islands are at greater risk of ill-health due to extreme heat events (Stone et al., 2010; Uejio et al., 2011). 11.3.2. Current Health Status Climate extremes may promote the transmission of certain infectious diseases and the vulnerability of populations to these diseases will depend on the baseline levels of pathogens and their vectors. In the United States, as one example, arboviral diseases such as dengue are rarely seen after flooding, compared with the experience in other parts of the Americas. The explanation lies in the scarcity of dengue (and other pathogenic viruses) circulating in the population, before the flooding (Keim, 2008). On the other hand, the high prevalence of HIV infection in many populations in Sub-Saharan Africa will tend to multiply the health risks of climate change, due to the interactions between chronic ill-health, poverty, extreme weather events and undernutrition (Ramin and McMichael, 2009). Chronic diseases such as diabetes and ischemic heart disease magnify the risk of death or severe illness associated with high ambient temperatures (Basu and Ostro, 2008; Sokolnicki et al., 2009). 11.3.3. Age and Gender Children, young people, and the elderly are at increased risk of climate-related injury and illness (Perera, 2008). For example, adverse effects of malaria, diarrhea, and undernutrition are presently concentrated amongst children, for reasons of physiological susceptibility (Michon et al., 2007). In principle, children are thought to be more vulnerable to heat-related illnesses, due to their small body mass to surface area ratio, but evidence of excess heat-related mortality in this age group is mixed (Basu and Ostro, 2008; Kovats and Hajat, 2008). Maternal antibodies acquired in utero provide some protection against dengue fever in the first year of life, but if infection does occur in infants it is more likely to provoke the severe hemorrhagic form of illness (Ranjit and Kissoon, 2011). Children are generally at greater risk when food supplies are restricted: households with children tend to have lower than average incomes, and food insecurity is associated with a range of adverse health outcomes amongst young children (Cook and Frank, 2008). Older people are at greater risk from storms, floods, heat-waves and other extreme events (Brunkard et al., 2008), in part because they tend to be less mobile than younger adults and so find it more difficult to avoid hazardous situations and also because they are more likely to live alone in some cultures. Older people are also more likely to suffer from health conditions that limit the body’s ability to respond to stressors such as heat and air pollution (Gamble et al., 2013). The relationship between gender and vulnerability is complex. Worldwide, mortality due to natural disasters, including droughts, floods and storms, is higher among women than men (World Health Organization, 2011). However there is variation regionally. In the United States, males are at greater risk of death following flooding (Jonkman and Kelman, 2005). A study of the health effects of flooding in Hunan province, China, also found an excess of flood deaths among males, often related to rural farming (Abuaku et al., 2009). In Canada’s Inuit population males are exposed to dangers associated with insecure sea ice, while females may be more vulnerable to the effects of diminished food supplies (Pearce et al., 2011). In the Paris 2003 heat wave, excess mortality was greater among females overall, but there were more excess deaths among men in the working age span (25 – 64) possibly due to differential exposures to heat in occupational settings (Fouillet et al., 2006). In Bangladesh, females are more affected than males by a range of climate hazards, due to differences in prevalence of poverty, undernutrition and exposure to water-logged environments (Neelormi et al., 2009). The effect of food insecurity on growth and development in childhood may be more damaging for girls than boys (Cook and Frank, 2008). Subject to Final Copyedit 8 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Pregnancy is a period of increased vulnerability to a wide range of environmental hazards, including extreme heat (Strand et al., 2012) and infectious diseases such as malaria, foodborne infections and influenza (Van Kerkhove et al., 2011). 11.3.4. Socioeconomic Status The poorest countries and regions are generally most susceptible to damage caused by climate extremes and climate variability (Malik et al., 2012), but wealthy countries are not immune, as shown by the deaths resulting from bushfires in Australia in 2009 (Teague et al., 2010). Also, rapid economic development may increase the risks of climate-related health issues. For instance, changes in Tibet, China including new roads and substantial in-migration may explain (along with above-average warming) the appearance and establishment in Lhasa of Culex pipiens, a mosquito capable of transmitting the West Nile virus (Liu et al., 2013b). A review of global trends in tropical cyclones 1970-2009 found that mortality risk at country-level depended most strongly on three factors: storm intensity, quality of governance, and levels of poverty (Peduzzi et al., 2012). Individuals and households most vulnerable to climate hazards tend to be those with relatively low socioeconomic status (Friel et al., 2008). A study of the impacts of flooding in Bangladesh found that household risk reduced with increases in both average income and number of income sources. Poorer households were not only more severely affected by flooding, but they took preventive action less often, and received assistance after flooding less frequently than did more affluent households (Brouwer et al., 2007). In many countries, race and ethnicity are powerful markers of health status and social disadvantage. Black Americans have been reported to be more vulnerable to heat-related deaths than other racial groups in the United States (Basu and Ostro, 2008). This may be due to a higher prevalence of chronic conditions such as over-weight and diabetes (Lutsey et al., 2010), financial circumstances (for instance, lower incomes may restrict access to air conditioning during heat-waves) (Ostro et al., 2010), or community-level characteristics such as higher local crime rates or disrupted social networks (Browning et al., 2006). Indigenous peoples who depend heavily on local resources, and live in parts of the world where climates are changing quickly, are generally at greater risk of economic losses and poor health. Studies of the Inuit people, for example, show that rapid warming of the Canadian Arctic is jeopardizing hunting and many other day-to-day activities, with implications for livelihoods and well-being (Ford, 2009). 11.3.5. Public Health and Other Infrastructure Populations that do not have access to good quality health care and essential public health services are more likely to be adversely affected by climate variability and climate change (Frumkin and McMichael, 2008). Harsh economic conditions in Europe since 2008 led to cutbacks in health services in some countries, followed by a resurgence of climate-sensitive infectious diseases including malaria (Karanikolos et al., 2013). The condition of the physical infrastructure that supports human settlements also influences health risks (this includes supply of power, provision of water for drinking and washing, waste management and sanitation: see Chapter 8). In Cuba, a country with a well-developed public health system, dengue fever has been a persistent problem in the larger cities, due in part to the lack of a constant supply of drinking water in many neighbourhoods (leading to people storing water in containers that are suitable breeding sites for the disease vector, A. aegypti) (Bulto et al., 2006). In New York, daily mortality spiked after a city-wide power failure in August 2003, due in part to increased exposure to heat (Weis, 2011). 11.3.6. Projections for Vulnerability Population growth is linked to climate change vulnerability. If nothing else changes, increasing numbers of people in locations that are already resource-poor and are affected by climate risks will magnify harmful impacts. Virtually all the projected growth in populations will occur in urban agglomerations, mostly in large, low latitude hot Subject to Final Copyedit 9 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 countries in which a high proportion of the workforce is deployed outdoors with little protection from heat. About 150 million people currently live in cities affected by chronic water shortages and by 2050, unless there are rapid improvements in urban environments, the number will rise to almost a billion (McDonald et al., 2011). Under a “business as usual” scenario with mid-range population growth, the OECD projects that about 1.4 billion people will be without access to basic sanitation in 2050 (OECD, 2012). The age structure of the population also has implications for vulnerability (see Figure 11-2). The proportion aged over 60, world-wide, is projected to increase from about 10% presently to about 32% by the end of the century (Lutz et al., 2008). The prevalence of overweight and obesity, which is associated with relatively poor heat tolerance, has increased almost everywhere in the last 20 years, and in many countries the trend continues upwards (Finucane et al., 2011). It has been pointed out that the Sahel region of Africa may be particularly vulnerable to climate change because it already suffers so much stress from population pressure, chronic drought, and governmental instability (Diffenbaugh and Giorgi, 2012; Potts and Henderson, 2012). [INSERT FIGURE 11-2 HERE Figure 11-2: Increasingly frequent heat extremes will combine with rapidly growing numbers of older people living in cities – who are particularly vulnerable to extreme heat. Countries are shaded according to the expected proportional increase in urban populations aged over 65 by the year 2050. Bar graphs show how frequently the maximum daily temperature that would have occurred only once in 20 years in the late 20th century is expected to occur in the mid-21st century, with lower numbers indicating more frequent events. Results are shown for 3 different “SRES” scenarios (Blue = B1; Green = A1B, Red = A2), as described in the IPCC Special Report on Emissions Scenarios, and based on 12 global climate models participating in the third phase of the Coupled Model Intercomparison Project (CMIP3). Coloured boxes show the range in which 50% of the model projections are contained, and whiskers show the maximum and minimum projections from all models. Source: World Health Organization and World Meteorological Organization, 2012.] Future trends in social and economic development are critically important to vulnerability. For instance, countries with a higher Human Development Index (HDI) (a composite of life expectancy, education, literacy and GDP per capita) are less affected by the floods, droughts and cyclones that take place (Patt et al., 2010). Therefore policies that boost health, education, and economic development should reduce future vulnerability. Overall, there have been substantial improvements in HDI in the last 30 years, but this has been accompanied by increasing inequalities between and within countries, and has come at the cost of high consumption of environmental resources (UNDP, 2011). 11.4. Direct Impacts of Climate and Weather on Health 11.4.1. Heat and Cold – Related Impacts Although there is ample evidence of the effects of weather and climate on health, there are few studies of the impacts of climate change itself. (An example: Bennett et al (2013) reported that the ratio of summer to winter deaths in Australia increased between 1968 and 2010, in association with rising annual average temperatures.) The issue is scale, since climate change is defined in decades. Robust studies require not only extremely long-term data series on climate and disease rates, but also information on other established or potential causative factors, coupled with statistical analysis to apportion changes in health states to the various contributing factors. Wherever risks are identified, health agencies are mandated to intervene immediately, biasing long-term analyses. Nevertheless, the connection between weather and health impacts is often sufficiently direct to permit strong inferences about cause and effect (Sauerborn and Ebi, 2012). Most notably, the association between hot days (commonly defined in terms of the percentiles of daily maximum temperature for a specified location) and increases in mortality is very robust (Honda et al., 2013). The IPCC Special Report on Extreme Events (SREX) concludes that it is very likely that there has been an overall decrease in the number of cold days and nights, and an overall increase in the number of warm days and nights, at the global scale. If there has been an increase in daily maximum temperatures, then it follows, in our view, that the number of heat-related deaths is likely to have also increased. For example, Christidis et al (2012) concluded that it is “extremely likely (probability greater than 95%)” that Subject to Final Copyedit 10 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 anthropogenic climate change at least quadrupled the risk of extreme summer heat events in Europe in the decade 1999-2008. The 2003 heat wave was one such record event: therefore the probability that particular heat wave can be attributed to climate change is 75% or more, and on this basis it is likely the excess mortality attributed to the heat wave (about 15,000 deaths in France alone (Fouillet et al., 2008)) was caused by anthropogenic climate change. The rise in minimum temperatures may have contributed to a decline in deaths associated with cold spells, however the influence of seasonal factors other than temperature on winter mortality suggests that the impacts on health of more frequent heat extremes greatly outweigh benefits of fewer cold days (Ebi and Mills, 2013; Kinney et al., 2012). Quantification, globally, remains highly uncertain, as there are few studies of the large developing country populations in the tropics, and these point to effects of heat, but not cold, on mortality (Hajat et al., 2010). There is also significant uncertainty over the degree of physiological, social or technological adaptation to increasing heat over long time periods. 11.4.1.1. Mechanisms The basic processes of human thermoregulation are well-understood. If the body temperature rises above 38°C (“heat exhaustion”), physical and cognitive functions are impaired; above 40.6°C (“heat stroke”), risks of organ damage, loss of consciousness, and death increase sharply. Detailed exposure-response relationships were described long ago (Wyndham, 1969), but the relationships in different community settings and for different age/sex groups are not yet well established. The early studies are supported by more recent experimental and field studies (Parsons, 2003; Ramsey and Bernard, 2000) and meta-analysis (Bouchama et al., 2007), that show significant effects of heat stress as body temperatures exceed 40°C, and heightened vulnerability in individuals with pre-existing disease. At high temperatures, displacement of blood to the surface of the body may lead to circulatory collapse. Indoor thermal conditions, including ventilation, humidity, radiation from walls or ceiling and the presence or absence of air-conditioning are important in determining whether adverse events occur, but these variables are seldom well- measured in epidemiological studies (Anderson et al., 2012). Biological mechanisms are less evident for other causes of death, such as suicide, that are sometimes related to high temperature (Kim et al., 2011; Likhvar et al., 2011; Page et al., 2007). Heat waves refer to a run of hot days; precisely how many days, and how high the temperatures must rise, are defined variously (Kinney et al., 2008). Some investigators have reported that mortality increases more during heat waves than would be anticipated solely on the basis of the short-term temperature mortality relationship (Anderson and Bell, 2011; D'Ippoliti et al., 2010), although the added effect is relatively small in some series, and most evident with prolonged heat waves (Gasparrini and Armstrong, 2011). Because heat waves are relatively infrequent compared with the total number of days with temperatures greater than the optimum for that location, the effects of heat waves are only a fraction of the total impact of heat on health. Some studies have shown larger effects of heat and heat waves earlier in the hot season (Anderson and Bell, 2011; Rocklov et al., 2011). This may be testament to the importance of acclimatisation and adaptive measures, or may result from a large group in the population that is more susceptible to heat early in the season (Rocklov et al., 2009; Rocklov et al., 2011). The extreme heat wave in Europe in 2003 led to numerous epidemiological studies. Reports from France (Fouillet et al., 2008) concluded that most of the extra deaths occurred in elderly people (80% of those who died were above 75 years). Questions were raised at the time as to why this event had such a devastating effect (Kosatsky, 2005). It is still not clear, but one contributing factor may have been the relatively mild influenza season the year before. Recent studies have found that when the previous year's winter mortality is low, the effect of summer heat is increased (Ha et al., 2011; Rocklov and Forsberg, 2009) because mild winters may leave a higher proportion of vulnerable people (Stafoggia et al., 2009). Most studies of heat have been in high-income countries, but there has been work recently in low- and middle-income countries, suggesting heterogeneity in vulnerability by age groups and socio-economic factors similar to that seen in higher-income settings (Bell et al., 2008; McMichael et al., 2008; Pudpong and Hajat, 2011). Subject to Final Copyedit 11 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Numerous studies of temperature-related morbidity, based on hospital admissions or emergency presentations, have reported increases in events due to cardio-vascular, respiratory and kidney diseases (Hansen et al., 2008; Knowlton et al., 2009; Lin and Chan, 2009) and the impact has been related to the duration and intensity of heat (Nitschke et al., 2011). There is evidence now that both average levels and variability in temperature are important influences on human health. The standard deviation of summer temperatures was associated with survival time in a US cohort study of persons aged over 65 years with chronic disease who were tracked from 1985-2006 (Zanobetti et al., 2012). Greater variability was associated with reduced survival. A study that modeled separately projected increases in temperature variability and average temperatures for six cities for 2070-2099 found that, with one exception, variability had an effect (increased deaths) over and above what was estimated from the rise in average temperatures (Gosling et al., 2009). Relevant to 11.5, rapid changes in temperature may also alter the balance between humans and parasites, increasing opportunities for new and resurgent diseases. The speed with which organisms adapt to changes in temperatures is, broadly speaking, a function of mass, and laboratory studies have shown that microbes respond more quickly to a highly variable climate than do their multi-cellular hosts (Raffel et al., 2012). Health risks during heat extremes are greater in people who are physically active (e.g. manual labourers). This has importance for recreational activity outdoors and it is relevant especially to the impacts of climate change on occupational health (11.6.2) (Ebi and Mills, 2013; Kjellstrom et al., 2009a). Heat also acts on human health through its effects, in conjunction with low rainfall, on fire risk. In Australia in 2009, record high temperatures, combined with long-term drought, caused fires of unprecedented intensity and 173 deaths from burns and injury (Teague et al., 2010). Smoke from forest fires has been linked elsewhere with increased mortality and morbidity (Analitis et al., 2012) (see 11.5.3.2). 11.4.1.2. Near-Term Future The climate change scenarios modeled by WG1 project rising temperatures and an increase in frequency and intensity of heat waves (2.6.1 and chapter 1 of this report) in the near-term future, defined as roughly midway through the 21st century, or the era of climate responsibility (see SPM). It is uncertain how much acclimatization may mitigate the effects on human health (Baccini et al., 2011; Bi and Parton, 2008; Hanna et al., 2011; Honda et al., 2013; Maloney and Forbes, 2011; Peng et al., 2011; Wilkinson et al., 2007; Wilkinson et al., 2007). In New York, it was estimated that acclimatization may reduce the impact of added summer heat in the 2050s by roughly a quarter (Knowlton et al., 2007). In Australia, the number of “dangerously hot” days, when core body temperatures may increase by ≥ 2ºC and outdoor activity is hazardous, is projected to rise from the current 4-6 days per year to 33-45 days per year by 2070 (with SRES A1FI) for non-acclimatized people. Among acclimatized people, an increase from 1-5 days per year to 5-14 days per year is expected (Hanna et al., 2011). For reasons given above, it is not clear whether winter mortality will decrease in a warmer, but more variable, climate (Ebi and Mills, 2013; Kinney et al., 2012). Overall, we conclude that the increase in heat-related mortality by mid-century will outweigh gains due to fewer cold periods, especially in tropical developing countries with limited adaptive capacities and large exposed populations (Wilkinson et al., 2007). A similar pattern has been projected for temperate zones. A study of three Quebec cities, based on SRES A2 and B2, extended to 2099, showed an increase in summer mortality that clearly outweighed a small reduction in autumn deaths, and only slight variations in winter and spring (Doyon et al., 2008). Another study in Brisbane, Australia, using years of life lost as the outcome, found the gains associated with fewer cold days were less than the losses caused by more hot days, when warming exceeded 2ºC. (Huang et al., 2012). A similar trend is reported in the United Kingdom (Health Protection Agency, 2012) and in New York City (Knowlton et al., 2007). Subject to Final Copyedit 12 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 11.4.2. Floods and Storms Floods are the most frequently occurring type of natural disaster (Guha-Sapir et al., 2011). In 2011, six of the 10 biggest natural disasters were flood events, when considered in terms of both number affected (112 million people) and number of deaths (3,140 people) (Guha-Sapir et al., 2011). Globally, the frequency of river flood events has been increasing, as well as economic losses, due to the expansion of population and property in flood plains (chapter 18). There is little information on health trends attributable to flooding, except for mortality and there are large differences in mortality risk between countries (UNISDR, 2011). Mortality from flooding and storm events is generally declining, but there is good evidence that mortality risks first increase with economic development before declining (De Haen and Hemrich, 2007; Kellenberg and Mobarak, 2008; Patt et al., 2010). For instance, migration to slums to coastal cities may increase population exposure at a greater pace than can be compensated for by mitigation measures (see chapter 10 on urban risks). Severe damaging floods in Australia in 2010-2011 and in the Northeastern United States in 2012 indicate that high-income countries may still be affected (Guha-Sapir et al., 2011). 11.4.2.1. Mechanisms Flooding and windstorms adversely affect health through drowning, injuries, hypothermia, infectious diseases (e.g. diarrhoeal disease, leptospirosis, vector-borne disease, cholera) (Jakubicka et al., 2010; Schnitzler et al., 2007). Since AR4, more evidence has emerged on the long term (months-years) implications of flooding for health. Flooding and storms may have profound effects on peoples’ mental health (Neria, 2012). The prevalence of mental health symptoms (psychological distress, anxiety and depression) was two to five times higher among individuals who reported flood water in the home compared to non-flooded individuals (2007 flood in England and Wales)(Paranjothy et al., 2011). In the United States, signs of hurricane-related mental illness were observed in a follow-up of New Orleans’ residents almost two years after Hurricane Katrina (Kessler et al., 2008). The attribution of deaths to flood events is complex; most reports of flood deaths include only immediate traumatic deaths, which means that the total mortality burden is under-reported (Health Protection Agency, 2012). There is some uncertainty as to whether flood events are associated with a longer-term (6-12 months) effect on mortality in the flooded population. No persisting effects were observed in a study in England and Wales (Milojevic et al., 2011), but longer-term increases in mortality were found in a rural population in Bangladesh (Milojevic et al., 2012). 11.4.2.2. Near-Term Future Under most climate change scenarios, it is expected that more frequent intense rainfall events will occur in most parts of the world in the future (IPCC, 2012). If this happens, floods in small catchments will be more frequent, but the consequence is uncertain in larger catchments (See Chapter 3). In terms of exposure, it is expected that more people will be exposed to floods in Asia, Africa and Central and South America (Chapter 3). Also, increases in intense tropical cyclones are likely in the late 21st century (WG1 Table SPM.1). It has been estimated conservatively that around 2.8 billion people were affected by floods between 1980-2009, with over 500,000 deaths (Doocy et al., 2013). On this basis we conclude it is very likely that health losses caused by storms and floods will increase this century if no adaptation measures are taken. What is not clear is how much of this projected increase can be attributed to climate change. Dasgupta et al. (2009) developed a spatially explicit mortality model for 84 developing countries and 577 coastal cities. They modelled 1:100 year storm-surge events, and assessed future impacts under climate change, accounting for sea level rise and a 10% increase in event intensity. In the 84 developing countries, an additional 52 million people and 30,000 km2 of land were projected to be affected by 2100. 11.4.3. Ultraviolet Radiation Ambient UV levels and maximum summertime day temperatures are related to the prevalence of non-melanoma skin cancers and cataracts in the eye. In one study in the United States, the number of cases of squamous cell Subject to Final Copyedit 13 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 carcinoma was 5.5% higher for every 1°C increment in average temperatures, and basal cell carcinoma was 2.9% more common with every 1°C increase. These values correspond to an increase in the effective UV dose of 2% for each 1°C (van der Leun et al., 2008). However, exposure to the sun has beneficial effects on synthesis of Vitamin D, with important consequences for health. Accordingly the balance of gains and losses due to increased UV exposures vary with location, intensity of exposure and other factors (such as diet) that influence Vitamin D levels (Lucas et al., 2013). Studies of stratospheric ozone recovery and climate change project that ultraviolet radiation levels at Earth’s surface will generally return to pre-1980 levels by mid-century, and may diminish further by 2100, although there is high uncertainty around the projections (Correa et al., 2013). On the other hand, higher temperatures in countries with temperate climates may result in an increase in the time which people spend outdoors (Belanger et al., 2009) and lead to additional UV-induced adverse effects. 11.5. Ecosystem-Mediated Impacts of Climate Change on Health Outcomes 11.5.1. Vector-Borne and Other Infectious Diseases Vector-borne diseases (VBDs) refer most commonly to infections transmitted by the bite of blood-sucking arthropods such as mosquitoes or ticks (Wong, Harris, Rodriguez-Galindo, & Johnson, 2013). These are some of the best-studied diseases associated with climate change, due to their widespread occurrence and sensitivity to climatic factors (Bangs et al., 2006; Bi et al., 2007; Halide and Ridd, 2008; Wu et al., 2009). Table 11-1 summarizes what is known about the influence of weather and climate on selected VBDs. [INSERT TABLE 11-1 HERE Table 11-1: The association between different climatic drivers and the global prevalence and geographic distribution of selected vector-borne diseases observed over the period 2008-2012. Among the vector borne diseases shown here, only dengue fever was associated with climate variables at both the global and local levels (high confidence), while malaria and hemorrhagic fever with renal syndrome showed a positive association at the local level (high confidence).] 11.5.1.1. Malaria Malaria is mainly caused by five distinct species of plasmodium parasite (Plasmodium falciparum, Plasmodium vivax, Plasmodium malariae, Plasmodium ovale, Plasmodium knowlesi), transmitted between individuals by Anopheline mosquitoes. In 2010 there were an estimated 216 million episodes of malaria worldwide, mostly amongst children under 5 years in the African Region (World Health Organization, 2010). The number of global malaria deaths was estimated to be 1,238,000 in 2010 (Murray et al., 2012). Worldwide, there have been significant advances made in malaria control in the last 20 years (Feachem, 2011). The influence of temperature on malaria development appears to be non-linear, and is vector-specific (Alonso et al., 2011). Increased variations in temperature, when the maximum is close to the upper limit for vector and pathogen, tend to reduce transmission, while increased variations of mean daily temperature near the minimum boundary increase transmission (Paaijmans et al., 2010). Analysis of environmental factors associated with the malaria vectors Anopheles gambiae and A. funestus in Kenya found that abundance, distribution, and disease transmission are affected in different ways by precipitation and temperature (Kelly-Hope et al., 2009). There are lag-times according to the lifecycle of the vector and the parasite: a study in central China reported that malaria incidence was related to the average monthly temperature, the average temperature of the previous two months, and the average rainfall of the current month (Zhang et al., 2012). More work has been done since AR4 to elucidate the role of local warming on malaria transmission in the East African highlands, but this is hampered by the lack of time series data on levels of drug resistance and intensity of vector control programs. Earlier research had failed to find a clear increase in temperatures accompanying increases in malaria transmission, but new studies with aggregated meteorological data over longer periods have confirmed increasing temperatures since 1979 (Omumbo et al., 2011; Stern et al., 2011). The strongly non-linear response to Subject to Final Copyedit 14 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 temperature means that even modest warming may drive large increases in transmission of malaria, if conditions are otherwise suitable (Alonso et al., 2011; Pascual et al., 2006). On the other hand, at relatively high temperatures modest warming may reduce the potential of malaria transmission (Lunde et al., 2013). One review (Chaves and Koenraadt, 2010) concluded that decadal temperature changes have played a role in changing malaria incidence in East Africa. But malaria is very sensitive also to socioeconomic factors and health interventions, and the generally more conducive climate conditions have been offset by more effective disease control activities. The incidence of malaria has reduced over much of East Africa (Stern et al., 2011) although increased variability in disease rates has been observed in some high altitude areas (Chaves et al., 2012). At the global level, economic development and control interventions have dominated changes in the extent and endemicity of malaria over the last 100 years (Gething et al., 2010). Although modest warming has facilitated malaria transmission (Alonso et al., 2011; Pascual et al., 2006), the proportion of the world's population affected by the disease has been reduced, largely due to control of P. vivax malaria in moderate climates with low transmission intensity. However, the burden of disease is still high and may actually be on the increase again, in some locations (World Health Organization, 2012). For instance, locally-transmitted malaria has re-emerged in Greece in association with economic hardship and cutbacks in government spending (Andriopoulos et al., 2013; Danis et al., 2011). 11.5.1.2. Dengue Fever Dengue is the most rapidly spreading mosquito-borne viral disease, showing a 30-fold increase in global incidence over the past 50 years (World Health Organization, 2013). Each year there occur about 390 million dengue infections worldwide, of which roughly 96 million manifest with symptoms (Bhatt et al., 2013). Three quarters of the people exposed to dengue are in the Asia-Pacific region, but many other regions are affected also. The first sustained transmission of dengue in Europe since the 1920s was reported in 2012 in Madeira, Portugal (Sousa et al., 2012). The disease is associated with climate on spatial (Beebe et al., 2009; Li et al., 2011; Russell et al., 2009), temporal (Descloux et al., 2012; Earnest et al., 2012; Gharbi et al., 2011; Herrera-Martinez and Rodriguez-Morales, 2010; Hii et al., 2009; Hsieh and Chen, 2009; Pham et al., 2011) and spatiotemporal (Chowell et al., 2008; Chowell et al., 2011; Lai, 2011) scales. The principal vectors for dengue, Aedes aegypti and Ae. albopictus, are climate-sensitive. Over the last two decades, climate conditions have become more suitable for albopictus in some areas (e.g. over central northwestern Europe) but less suitable elsewhere (e.g. over southern Spain) (Caminade et al., 2012). Distribution of Ae. albopictus in northwestern China is highly correlated with annual temperature and precipitation (Wu et al., 2011). Temperature, humidity and rainfall are positively associated with dengue incidence in Guangzhou, China, and wind velocity is inversely associated with rates of the disease (Li et al., 2011; Lu and Lin, 2009). Several studies in Taiwan reported that typhoons remain an important factor affecting vector population and dengue fever (Hsieh and Chen, 2009; Lai, 2011). Typhoons result in extreme rainfall, high humidity and water pooling, and may generate fresh mosquito breeding sites. A study in Dhaka, Bangladesh reported increased rates of admissions to hospital due to dengue with both high and low river levels (Hashizume and Dewan, 2012). In some circumstances, it is apparent that heavy precipitation favors the spread of dengue fever, but drought can also be a cause if households store water in containers that provide suitable mosquito breeding sites (Beebe et al., 2009; Padmanabha et al., 2010). _____ START BOX 11-2 HERE _____ Box 11-2. Case Study: An Intervention to Control Dengue Fever Seasonality in dengue transmission is well established in many parts of the world, and transmission occurs mostly during the wettest months of the year (Chadee et al., 2007; Gubler and Kuno, 1997). Figure 11-3 shows that about 80% of dengue fever cases in Trinidad were recorded during the wet season, a period when the Ae. aegypti mosquito population density was four to nine times higher than the dengue transmission threshold (Macdonald, 1956). This led to a control program that concentrated on reducing the mosquito population before the onset of the rains, by application of insecticides (temephos) into the water drums that serve as primary breeding sites of Ae. aegypti in the Subject to Final Copyedit 15 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Caribbean. The one-off treatment effectively controlled the mosquito populations for almost 12 weeks after which the numbers reverted to levels observed in the untreated control areas. [INSERT FIGURE 11-3 HERE Figure 11-3: a) Rainfall, temperature, Breteau index (number of water containers with Ae. aegypti larvae per 100 houses), and dengue fever cases, Trinidad (2002-2004). Rainfall was found to be significantly correlated with an increase in the Ae. aegypti population and dengue fever incidence, with a clearly defined “dengue season” between June and November over two years of the study. Source: (Chadee et al., 2007). b) Efficacy of pre-seasonal treatment with temephos on Ae. aegypti ovitrap egg counts in Curepe (treatment) and St. Joseph (control), Trinidad (2003). Evidence of the efficacy of the pre-seasonal larval control through focal treatment of Ae. aegypti population is provided. Treatment at the onset of the rainy season can effectively prevent the rapid increase in Ae. aegypti populations and therefore suppress the onset of dengue transmission. Source: Chadee, 2009.] Climate scenarios that extend to 2071-2100 project changes in the intensity and frequency of rainfall events in the Caribbean (Campbell et al., 2011). In these scenarios, there is greater variability in rainfall patterns during November to January, with the northern Caribbean region receiving more rainfall than in the southern Caribbean (Campbell et al., 2011). There may be water shortages during drought periods, and flooding after episodes of heavy rainfall, both of which affect the breeding habitats of Ae. aegypti and Ae. albopictus. Vector control strategies will need to be planned and managed astutely to systematically reduce mosquito populations. _____ END BOX 11-2 HERE _____ 11.5.1.3. Tick-Borne Diseases Tick-borne encephalitis (TBE) is caused by tick-borne encephalitis virus, and is endemic in temperate regions of Europe and Asia. Lyme disease is an acute infectious disease caused by the spirochaete bacteria Borrelia burgdorferi and is reported in Europe, the USA and Canada. Borrelia is transmitted to humans by the bite of infected ticks belonging to a few species of the genus Ixodes ("hard ticks"). Many studies have reported associations between climate and tick-borne diseases (Andreassen et al., 2012; Estrada-Peña et al., 2012; Jaenson et al., 2012; Lukan et al., 2010; Okuthe and Buyu, 2006; Tokarevich et al., 2011). In North America, there is good evidence of northward expansion of the distribution of the tick vector (Ixodes scapularis) in the period 1996-2004 based on an analysis of active and passive surveillance data (Ogden et al., 2010). However, there is no evidence so far of any associated changes in the distribution in North America of human cases of tick-borne diseases. There was a marked rise in TBE cases from the 1970s in central and eastern Europe. Spring-time daily maximum temperatures rose in the late 1980s, sufficient to encourage transmission of the TBE virus. For instance, in the Czech Republic, between 1970 and 2008, there were signs of lengthening transmission season and higher altitudinal range in association with warming (Kriz et al., 2012). However variations in illness rates across the region demonstrate that climate change alone cannot explain the increase. Socioeconomic changes (including changes in agriculture and recreational activities) have affected patterns of disease in Europe (Randolph, 2010; Sumilo et al., 2008). The complex ecology of tick-borne diseases such as Lyme disease and TBE make it difficult to attribute particular changes in disease frequency and distribution to specific environmental factors such as climate (Gray et al., 2009). 11.5.1.4. Other Vector-Borne Diseases Hemorrhagic fever with renal syndrome (HFRS) is a zoonosis caused by the Hanta virus, and leads to approximately 200,000 hospitalized cases each year. The incidence of this disease has been associated with temperature, precipitation, and relative humidity (Fang et al., 2010; Liu et al., 2011; Pettersson et al., 2008). Plague, one of the oldest diseases known to humanity, persists in many parts of the world. Outbreaks have been linked to seasonal and inter-annual variability in climate (Holt et al., 2009; MacMillan et al., 2012; Nakazawa et al., 2007; Stenseth et al., Subject to Final Copyedit 16 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 2006; Xu et al., 2011). Chikungunya fever is a climate-sensitive mosquito-transmitted viral disease (Anyamba et al., 2012), first identified in Africa, now present also in Asia, and the disease has recently emerged in parts of Europe (Angelini et al., 2008). The incidence in China of Japanese encephalitis, another mosquito-borne viral disease, is correlated with temperature and rainfall, especially during the warmer months of the year (Bai et al., 2013). In West Africa, outbreaks of Rift Valley Fever, an acute viral disease affecting humans and domestic animals, are linked to within-season variability in rainfall (Caminade et al., 2011). 11.5.1.5. Near-Term Future Using the A1B climate change scenario, Béguin et al. (2011) projected the population at risk of malaria to 2030 and 2050. With GDP per capita held constant at 2010 values, the model projected 5.2 billion people at risk in 2050, out of a predicted global population of 8.5 billion. Keeping climate constant, and assuming strong economic growth allied with social development (“best case”), the model projected 1.74 billion people at risk (approximately half the present number at risk) in 2050. Factoring in climate change would increase the “best case” estimate of the number of people at risk of malaria in 2050 to 1.95 billion, which is 200 million more than if disease control efforts were not opposed by higher temperatures and shifts in rainfall patterns. There are no recent studies that project the return of established malaria to North America or Europe, where it was once prevalent. However suitable vectors for P. vivax malaria abound in these parts of the world, and recent experience in southern Europe demonstrates how rapidly the disease may re-appear if health services falter (Bonovas and Nikolopoulos, 2012). A systematic review of research on the distribution of dengue and possible influence of climate change (Van Kleef et al., 2010) concluded that the area of the planet that was climatically suitable for dengue would increase under most scenarios, but it was not possible to project the impact on disease incidence. Åström (Astrom et al., 2012) estimated the population at risk out to the year 2050. The study was based on routine disease reports, surveys, population projections, estimates of GDP growth and the A1B scenario for climate change. Assuming high GDP growth that benefits all populations, the number exposed to dengue in 2050 falls to 4.46 billion, i.e. the adverse effects of climate change are balanced by the beneficial outcomes of development. This study considered only the margins of the geographic distribution of dengue (where economic development has its strongest effect) and did not examine changes in intensity of transmission in areas where the disease is already established. Kearny (2009) used biophysical models to examine the potential extension of vector range in Australia. He predicted that climate change would increase habitat suitability throughout much of Australia. Changes in water storage as a response to a drier climate may be an indirect pathway, through which climate change affects mosquito breeding (Beebe et al., 2009). 11.5.2. Food- and Water-Borne Infections Human exposure to climate-sensitive pathogens occurs by ingestion of contaminated water or food, incidental ingestion during swimming or by direct contact with eyes, ears or open wounds. Pathogens in water may be zoonotic in origin, concentrated by bivalve shellfish (e.g., oysters) or deposited on irrigated food crops. Pathogens of concern include enteric organisms that are transmitted by the fecal oral route and also bacteria and protozoa that occur naturally in aquatic systems. Climate may act directly by influencing growth, survival, persistence, transmission or virulence of pathogens; indirect influences include climate-related perturbations in local ecosystems or the habitat of species that act as zoonotic reservoirs. 11.5.2.1. Vibrios Vibrio is a genus of native marine bacteria that includes a number of human pathogens, most notably V. cholerae which causes cholera. Cholera may be transmitted by drinking water or by environmental exposure in seawater and Subject to Final Copyedit 17 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 seafood; other Vibrio species are solely linked to seawater and shellfish. These include V. parahaemolyticus and V. vulnificus, with V. alginolyticus emerging in importance (Weis, 2011). Risk of infection is influenced by temperature, precipitation and accompanying changes in salinity due to freshwater run-off, addition of organic carbon or other nutrients or changes in pH. These factors all affect the spatial and temporal range of the organism and also influence exposure routes (e.g. direct contact or via seafood). In countries with endemic cholera, there appears to be a robust relationship between temperature and the disease (Islam, 2009; Paz, 2009; Reyburn et al., 2011). In addition, heavy rainfall promotes the transmission of pathogens when there is not secure disposal of fecal waste. An unequivocal positive relationship between Vibrio numbers and SST in the North Sea has been established by DNA analyses of formalin-fixed samples collected over a 44 year period (Vezzulli et al., 2012). Cholera outbreaks have been linked to variations in temperature and rainfall, and other variables including sea and river levels, sea chlorophyll and cyanobacteria contents, Indian Ocean Dipole (IOD) and El Niño–Southern Oscillation (ENSO) events (Bompangue et al., 2011; Constantin de Magny et al., 2008; Hashizume, 2008; Reyburn et al., 2011; Rinaldo et al., 2012). 11.5.2.2. Other Parasites, Bacteria, and Viruses Rates of diarrhea have been associated with high temperatures (Kolstad and Johansson, 2011). Mostly, however, the specific causes of the diarrheal illness are not known, nor the mechanism for the association with temperature. Exceptions include Salmonella and Campylobacter, among the most common zoonotic food and waterborne bacterial pathogens worldwide, which both show distinct seasonality in infection and higher disease rates at warmer temperatures. The association between climate (especially temperature) and non-outbreak (‘sporadic’) cases of salmonellosis may, in part, explain seasonal and latitudinal trends in diarrhea (Lake, 2009). Among the enteric viruses, there are distinct seasonal patterns in infection that can be related indirectly to temperature. Enterovirus infections in the U.S. peak in summer and fall months (Khetsuriani et al., 2006). After controlling for seasonality and interannual variations, hand, foot and mouth disease (caused by coxsackievirus A16 and enterovirus 71), shows a linear relationship with temperature in Singapore, with a rapid rise in incidence when the temperature exceeds 32°C (Hii et al., 2011). However, it is not clear what the underlying driver is and if temperature is confounded by other seasonal factors. Temperature is directly linked with risk of enteric disease in Arctic communities, since melt of the permafrost hastens transport of sewage (which is often captured in shallow lagoons) into groundwater, drinking water sources or other surface waters (Martin et al., 2007). Additionally, thawing may damage drinking water intake systems (for those communities with such infrastructure) (Hess, 2008). Rainfall has also been associated with enteric infections. Bacterial pathogens are more likely to grow on produce crops (e.g. lettuce) in simulations of warmer conditions (Liu et al., 2013a), and become attached to leafy crops under conditions of both flooding and drought (Ge et al., 2012). This latter pattern is reflected in patterns of illness (Bandyopadhyay et al., 2012). Higher concentrations of enteric viruses have been reported frequently in drinking water and recreational water following heavy rainfall (Delpla et al., 2009). Worldwide, rotavirus infections caused about 450,000 deaths in children under 5 years old in 2008 (Tate et al., 2012). There are seasonal peaks in the number of cases in temperate and subtropical regions but less distinct patterns are seen within 10° latitude of the equator (Cook et al., 1990). Variations in the timing of peak outbreaks between countries or regions (Atchison et al., 2010; Turcios et al., 2006) and variations with time in the same country (Dey et al., 2010) have been attributed to fluctuations in the number and seasonality of births (Pitzer et al., 2009; Pitzer et al., 2011). While vaccination against rotavirus is expected to reduce the total burden of disease, it may also increase seasonal variation (Pitzer et al., 2011; Tate et al., 2009). Harmful algal blooms can be formed by (i) dinoflagellates that cause outbreaks of paralytic shellfish poisoning, ciguatera fish poisoning and neurotoxic shellfish poisoning; (ii) cyanobacteria that produce toxins causing liver, neurological, digestive and skin diseases; and (iii) diatoms that can produce domoic acid, a potent neurotoxin which is bioaccumulated in shellfish and finfish (Erdner et al., 2008). Increasing temperatures promote bloom formation in Subject to Final Copyedit 18 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 both fresh water (Paerl et al., 2011), and marine environments (Marques et al., 2010). (See Chapter 5.) Increasing temperature favoured growth of toxic over non-toxic strains of Microcystis in lakes in the USA (Davis et al., 2009). Projections of toxin-producing blooms in Puget Sound using an A1B scenario suggest that by the end of the century the “at risk” period may begin 2 months earlier and last up to 1 month longer than at present (Moore et al., 2011). 11.5.2.3. Near-Term Future Kolstad and Johansson (2011) projected an increase of 8-11% in the risk of diarrhea in the tropics and subtropics in 2039 due to climate change, using the A1B scenario and 19 coupled atmosphere-ocean climate models from CMIP3. This study did not account for future changes in economic growth and social development. Application of down- scaled climate change models showed that overflows of sewage into Chicago’s watersheds would increase by 50- 120% by 2100, as a result of more frequent and intense rainfall (Patz et al., 2008). In Botswana, if hot, dry conditions begin earlier in the year, and are prolonged, as projected by down-scaled climate scenarios, the present dry season peak in diarrhoeal disease may be amplified (Alexander et al., 2013). However the same analysis projected that incidence of diarrhoeal disease in the wet season would decline. Zhou et al. (2008) studied the effect of climate on transmission of schistosomiasis due to S. japonicum in China. They concluded that an additional 784 thousand km2 would become suitable for schistosomiasis transmission in China by 2050, as the mid-winter freezing line moves northwards (Figure 11-4). [INSERT FIGURE 11-4 HERE Figure 11-4: Effect of rising temperatures on the area in which transmission of Schistosomiasis japonica may occur. Green area denotes the range of schistosomiasis in China in 2000. The blue area shows the additional area suitable for disease transmission in 2050. Based on a biology-driven model including parasite (Schistosoma japonicum) and snail intermediate host (Oncomelania hupensis) and assuming average temperatures in China in mid-winter (January) increase by 1.6°C in 2050, compared with 2000. Adapted from Zhou et al., 2008.] Mangal et al. (2008) constructed a mechanistic model of the transmission cycle of another species, S. mansoni, and reported a peak in the worm burden in humans at an ambient temperature of 30°C, falling sharply as temperature rises to 35°C. The authors attribute this to the increasing mortality of both the snails and the water-borne intermediate forms of the parasite, and noted that worm burden is not directly linked to the prevalence of schistosomiasis. 11.5.3. Air Quality Nearly all the non-CO2 climate-altering pollutants (CAPs – see Chapters 7 and 8 of WGI) are health damaging, either directly or by contributing to secondary pollutants in the atmosphere. Thus, like the ocean acidification and ecosystem/ agriculture fertilization impacts of CO2, the other CAPs have non-climate-mediated impacts, particularly on health. Although not reviewed in detail in this assessment, the health impacts of non-CO2 CAPs are substantial globally. See Box 11-3. _____ START BOX 11-3 HERE _____ Box 11-3. Health and Economic Impacts of Climate-Altering Pollutants (CAPs) Other than CO2 Although other estimates of the global health impacts of human exposures to particle and ozone pollution have been published in recent years (e.g. (UNEP, 2011)), the most comprehensive was the Comparative Risk Assessment carried out as part of the 2010 Global Burden of Disease Project (Lim et al., 2012). It found that the combined health impact of the household exposures to particle air pollution from poor combustion of solid cooking fuels, plus general ambient pollution, was about 6.8 million premature deaths annually, with about 5 percent overlapping, i.e., coming from the contribution to general ambient pollution of household fuels. It also found that about 150 thousand premature deaths could be attributed to ambient ozone pollution. Put into terms of disability-adjusted life years Subject to Final Copyedit 19 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 (DALYs), particle air pollution was responsible for about 190 million lost DALYs in 2010, or about 7.6% of all DALYs lost. This burden puts particle air pollution among the largest risk factors globally, far higher than any other environmental risk and rivaling or exceeding all of the five dozen risk factors examined, including malnutrition, smoking, high blood pressure, and alcohol. The economic impact of this burden is difficult to assess as evaluation methods vary dramatically in the literature. Most in the health field prefer to consider some version of a lost healthy life year as the best metric although the economics literature often uses willingness to pay for avoiding a lost life (Jamison et al., 2006). Another difficulty is that any valuation technique that weights the economic loss according to local incomes per capita will value health effects in rich countries more than in poor countries, which would seem to violate some of the premises of a global assessment; see WGIII Chpt 3 for more discussion. Here, however, we will use the mean global income per capita (~ USD 10,000 in 2010) to scope out the scale of the impact globally without attempting to be specific by country or region. The WHO CHOICE approach for evaluating what should be spent on health interventions indicates that one annual per capita income per DALY is a reasonable upper bound (World Health Organization, 2009a). This would imply that the total lost economic value from global climate-altering pollutants in the form of particles is roughly USD 1.9 trillion, in the sense that the world ought to be willing to pay this much to reduce it. This is about 2.7% of the global economy (approximately USD 70 trillion in 2010). On the one hand, this shows that global atmospheric pollution already has a major impact on the health and economic well-being of humanity today, due mainly to the direct effects rather than those mediated through climate. If CO2 is not controlled and climate change continues to intensify while air pollutant controls become more stringent, the climate impacts will become more prominent. The quite different time scales for the two types of impacts make comparisons difficult, however. Air pollution reductions do not always promote the twin goals of protecting health and climate but can pose trade- offs. All particles are dangerous for health, for example, but some are cooling, such as sulfates, and some warming, such as black carbon (Smith et al., 2009). Indeed elimination of all anthropogenic particles in the atmosphere, a major success for health, would have only a minor net impact on climate (WG1, Fig TS-6). As discussed in the co- benefits section below (11.9), there are nevertheless specific actions that will work toward both goals. _____ END BOX 11-3 HERE _____ Although there is a large body of literature on the health effects of particulate air pollution (see Box 11-3), WGI indicates that there is little evidence that climate change, per se, will affect long-term particle levels in a consistent way (WG1, 11.3.5; Annex II). Thus, we focus here on chronic ozone exposures, which are found by WGI to be enhanced in some, but not all, scenarios of future climate change (WG1, TS.5.4.8). 11.5.3.1. Long-Term Outdoor Ozone Exposures Tropospheric ozone is formed through photochemical reactions that involve nitrogen oxides (NOx), carbon monoxide (CO), methane (CH4), and volatile organic compounds (VOCs) in the presence of sunlight and elevated temperatures (WG1, Ch 8). Therefore, if temperatures rise, many air pollution models (Chang et al., 2010; Ebi and McGregor, 2008; Polvani et al., 2011; Tsai et al., 2008) project increased ozone production especially within and surrounding urban areas (Hesterberg et al., 2009). Enhanced temperature also accelerates destruction of ozone, and the net direct impact of climate change on ozone concentrations worldwide is thought to be a reduction (WG1 TS.5.4.8). Some WG1 (TS.5.4.8) scenarios, however, indicate tropospheric ozone may rise from additional methane emissions stimulated by climate change. Models also show that local variations can have a different sign to the global one (Selin et al., 2009). Even small increases in atmospheric concentrations of ground-level ozone may affect health (Bell et al., 2006; Ebi and McGregor, 2008; Jerrett et al., 2009). For instance, Bell et al. (2006) found that levels that meet the US EPA 8- Subject to Final Copyedit 20 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 hour regulation (0.08 ppm over 8 hours) were associated with increased risk of premature mortality. There is a lack of association between ozone and premature mortality only at very low concentrations (from 0 to ~10 ppb) but the association becomes positive and approximately linear at higher concentrations (Bell et al., 2006; Ebi and McGregor, 2008; Jerrett et al., 2009). In an analysis of 66 United States cities with 18 years of follow-up (1982- 2000), tropospheric ozone levels were found to be significantly associated with cardiopulmonary mortality (Smith et al., 2009). See also the global review by WHO, which includes data from developing countries (World Health Organization, 2006). 11.5.3.2. Acute Air Pollution Episodes Wildfires, which occur more commonly following heat waves and drought, release particulate matter and other toxic substances that may affect large numbers of people for days to months (Finlay et al., 2012; Handmer et al., 2012). During a fire near Denver (USA) in June 2009, 1-hour concentrations of PM10 and PM2.5 reached 370 µg/m3 and 200 µg/m3, and 24-hour average concentrations reached 91 µg/m3 and 44 µg/m3, compared to the 24-hour WHO air quality guidelines for these pollutants of 50 µg/m3 and 25 µg/m3, respectively (Vedal and Dutton, 2006). One study of worldwide premature mortality attributable to air pollution from forest fires estimated there were 339,000 deaths per year (range 260,000 to 600,000) (Johnston et al., 2012). The regions most affected are Sub-Saharan Africa and Southeast Asia (Johnston et al., 2012). Extremely high levels of PM10 were observed in Moscow due to forest fires caused by a heat wave in 2010. Daily mean temperatures in Moscow exceeded the respective long-term averages by 5°С or more for 45 days. Ten new temperature records were established in July and nine in August, based on measurements since 1885, and an anti- cyclone in the Moscow region prevented dispersion of air pollutants. The highest 24-h pollution levels recorded in Moscow during these conditions were between 430 and 900 µg/m 3 PM10 most days, but occasionally reached 1500 µg/m3. The highest 24-h CO concentration was 30 mg/m3 compared to the WHO AQG of 7 µg/m3, and the levels of formaldehyde, ethyl benzene, benzene, toluene and styrene were also increased (State Environmental Institution “Mosecomonitoring”, 2010). There may be an interaction of tropospheric ozone and heat waves. Dear et al. (2005) modeled the daily mortality due to heat and exposure to ozone during the European summer heat wave of 2003 and found that possibly 50% of the deaths could have been associated with ozone exposure rather than the heat itself. 11.5.3.3. Aeroallergens Allergic diseases are common and are climate-sensitive. Warmer conditions generally favour the production and release of air-borne allergens (such as fungi and lower plant spores and pollen) and, consequently, there may be an effect on asthma and other allergic respiratory diseases, such as asthma and allergic rhinitis, as well as effects on conjunctivitis and dermatitis (Beggs, 2010). Children are particularly susceptible to most allergic diseases (Schmier and Ebi, 2009). Increased release of allergens may be amplified if higher CO2 levels stimulate plant growth. Visual monitoring and experiments have shown that increases in air temperature cause earlier flowering of prairie tallgrass (Sherry et al., 2007). Droughts and high winds may produce windborne dust and other atmospheric materials, which contain pollen and spores, and transport these allergens to new regions. Studies have shown that increasing concentrations of grass pollen lead to more frequent ambulance calls due to asthma symptoms, with a time lag of 3-5 days (Heguy et al., 2008). Pollen levels have also been linked to hospital visits with rhinitis symptoms (Breton et al., 2006). A cross-sectional study in the three climatic regions of Spain documented a positive correlation between the rate of child eczema and humidity, and negative correlation between child eczema and air temperature or the number of sunshine hours (Suarez-Varela et al., 2008). Subject to Final Copyedit 21 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 11.5.3.4. Near-Term Future It is projected by WGI that climate change could affect future air quality, including levels of photochemical oxidants and, with much less certainty, fine particles (PM2.5). If this occurs, there will be consequences for human health (Bell et al., 2007; Chang et al., 2012; Dong et al., 2011; Lepeule et al., 2012; Meister et al., 2012; West et al., 2013). High temperatures may also magnify the effects of ozone (Jackson et al., 2010; Ren et al., 2008). Increasing urbanization, use of solid biomass fuels, and industrial development in the absence of emission controls could also lead to increases in ozone chemical precursors (Selin et al., 2009; Wilkinson et al., 2009). Most post-2006 studies on the projected impacts of future climate change on air pollution-related morbidity and mortality have focused on ozone in Europe, the U.S. and Canada (see Table 11-S1; (Bell et al., 2007; Selin et al., 2009; Tagaris et al., 2009)). Projections are rare for other areas of the world, notably the developing countries where air pollution is presently a serious problem and is expected to worsen unless controls are strengthened. Higher temperatures may magnify the effects of air pollutants like ozone, although estimates of the size of this effect vary (Jackson et al., 2010; Ren et al., 2008). In general, all-cause mortality related to ozone is expected to increase in the US and Canada (Bell et al., 2007; Cheng et al., 2011; Jackson et al., 2010; Tagaris et al., 2009). Under a scenario in which present air quality legislation is rolled out everywhere, premature deaths due to ozone would be wound back in Africa, South Asia and East Asia. Under a maximum feasible CO2 reduction scenario related to A2, it is projected that 460,000 premature ozone-related deaths could be avoided in 2030, mostly in South Asia (West et al., 2007). All-cause mortality, however, is not the best metric for comparing air pollution health impacts across regions; given that background disease conditions vary so widely. HIV deaths and malaria deaths, which are prominent in Sub-Saharan Africa, for example, are not expected to increase from air pollution exposures in the same way as deaths from cardiovascular disease that dominate other regions. A study that investigated regional air quality in the United States in 2050, using a down-scaled climate model (Goddard Institute for Space Studies, Global Climate Model), concluded there would be about 4000 additional annual premature deaths due to increased exposures to PM2.5 (Tagaris et al., 2009). Air pollutant-related mortality increases are also projected for Canada, but in this case they are largely driven by the effects of ozone (Cheng et al., 2011). On the basis of the relation of asthma to air quality in the last decade (1999-2010), Thompson et al. (2012) anticipate that the prevalence of asthma in South Africa will increase substantially by 2050. Sheffield et al. (2011), applying the SRES A2 scenario, projected a median 7.3% increase in summer ozone-related asthma emergency department visits for children (0-17 years) across New York City by the 2020s compared to the 1990s. 11.6. Health Impacts Heavily Mediated through Human Institutions 11.6.1. Nutrition Nutrition is a function of agricultural production (net of post-harvest wastes and storage losses), socioeconomic factors, such as food prices and access, and human diseases, especially those which affect appetite, nutrient absorption and catabolism (Black et al., 2008; Lloyd et al., 2011). All three may be influenced by climate but only agricultural production has been modeled in a climate impacts framework. Here we use the terms undernutrition, which is a health outcome, and undernourishment, which reflects national (post-trade) calories available for human consumption, and is expressed as estimated percent of the population receiving ‘insufficient’ calories. We do not use the term “malnutrition,” as it includes overnutrition, which is not considered here (except under co-benefits in 11.9). Undernutrition can be chronic, leading to stunting (low height for age) or acute, leading to wasting (low weight for height); underweight (low weight for age) is a combination of chronic and acute undernutrition. 11.6.1.1. Mechanisms The processes through which climate change can affect human nutrition are complex (see chapter 7.2.2). Higher temperatures and changes in precipitation may reduce both the quantity and quality of food harvested (e.g. (Battisti Subject to Final Copyedit 22 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and Naylor, 2009)). Lobell et al. (2011b) showed for African maize that for each degree above 30°C, yields decreased by 1% under optimal rainfall conditions and by 1.7 % under drought conditions. From their systematic review of more than a thousand studies, Knox et al. (2012) drew the conclusion that “climate change is a threat to crop productivity in areas that are already food insecure.” Grace et al. (2012) modeled the relationship between climate variables (temperature and precipitation), food production and availability as well as child stunting in Kenya. The authors conclude that climate change will increase the proportion of stunted children in countries such as Kenya that are dependent on rain-fed agriculture, unless there are substantial adaptation efforts, such as investment in education and agricultural technology. Rising temperatures may also affect food security through the impact of heat on productivity of farmers (see 11.6.2). The magnitude of detected and predicted decline in land-based agricultural production due to increasing temperatures and changes in rainfall must be put in perspective to other changes, such as increase in harvests due to improved farming knowledge and technology, the amount of food fed to livestock, used for biofuels, consumed beyond baseline needs by the overnourished or wasted in other ways (Foley et al., 2011). There is good evidence that local food price increases have negative effects on food consumption, and therefore on health (Green et al., 2013). Against this background, the global food price fluctuates, though with a recently rising trend. While the main driver is higher energy costs, amplified by speculation (Piesse and Thirtle, 2009), there is growing evidence (Auffhammer, 2011) that extreme weather events, especially floods, droughts (Williams and Funk, 2011) and heat waves, may have contributed to higher prices. All else being equal, higher prices increase the number of malnourished people. See Chapter 7 for a more detailed discussion of the impact of climate change on food production. 11.6.1.2. Near-Term Future Since AR4 at least four studies have been published which project the effect of climate change on undernourishment and undernutrition. Nelson et al. (2010; 2009) conducted two studies using a crop simulation model (DSSAT) and a global agricultural trade model (IMPACT 2009) to estimate crop production (with and without CO2 enrichment), calorie availability, child underweight, and adaptation costs. The first study (Nelson et al., 2009) was carried out under the A2 emission scenario, using two GCMs (NCAR, CSIRO) and relative to a ‘no climate change’ future. The authors found that yields of most important crops would decline in developing countries by 2050, that per capita calorie availability would drop below levels that applied in the year 2000, and that child underweight would be ~20% higher (in the absence of carbon enrichment effects). That is, about 25 million children would be affected. (See Table 11-2). Of note, the underweight estimates do not account for possible improvements in socioeconomic conditions between 2000 and 2050. However it was estimated that substantial improvements would be necessary to counteract the effects of climate change. These included a 60% increase in yield growth (all crops) over baseline, 30% faster growth in animal numbers, and a 25% increase in the rate of expansion of irrigated areas. The second study by Nelson et al. used a wider range of socioeconomic and climate scenarios but health impacts were similar to the first study. Estimates of improved socioeconomic conditions were insufficient to fully offset the potential impacts of climate change: child underweight was estimated to be ~10% higher with climate change compared to a future without climate change. [INSERT TABLE 11-2 HERE Table 11-2: Number of under-nourished children less than 5 years of age (in millions) in 2000 and 2050, using the NCAR (National Center for Atmospheric Research) climate model (and the A2 scenario from AR4). Results assume no effect of heat on farmers’ productivity, and no CO2 fertilization benefits. Adapted from (Nelson et al., 2009).] Lloyd et al. (2011) built a model for estimating future stunting driven by two principle inputs: estimates of undernourishment (i.e. ‘food-related’ causes of stunting) and socioeconomic conditions (i.e. ‘non-food related’ causes of stunting). The former were based on calorie availability estimates from Nelson et al. (2009), and the latter on GDP per capita projections and estimates of the Gini index for income distribution. They estimated that by 2050, under A2 emissions with moderate to high economic growth and compared to a future without climate change, there Subject to Final Copyedit 23 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 may be a relative increase of severe stunting of 31% to 55% across regions of Sub-Saharan Africa and 61% in South Asia. It should be noted here that severe stunting carries 3 to 4 times the mortality risk of moderate stunting. In a future without climate change, undernutrition was projected to decline, leading the authors to conclude that climate change would hold back efforts to reduce child undernutrition in the most severely affected parts of the world, even after accounting for the potential benefits of economic growth. In addition to global studies, regional projections of the impacts of climate change on undernutrition have also been carried out since AR4. Grace et al. (2012) modeled the relationship between climate variables (temperature and precipitation), food production and availability as well as child stunting in Kenya. The authors conclude that climate change will increase the proportion of stunted children in countries such as Kenya that are dependent on rain-fed agriculture, unless there are substantial adaptation efforts, such as investment in education and agricultural technology. Similarly, Jankowska et al. (Jankowska et al., 2012) included climate, livelihood and health variables (stunting and underweight). The authors identified a link between type of livelihood and risk of undernutrition, and climate and stunting. Applying the model to Mali, the authors projected impacts to 2025 and estimated that nearly 6 million people may experience undernutrition due to changes in climate, livelihood and demography; three-quarter to one million of this number will be children under five. In summary, we conclude that climate change will have a substantial negative impact on (i) per capita calorie availability, (ii) childhood undernutrition, particularly stunting and (iii) on undernutrition–related child deaths and DALYs lost in developing countries (high confidence). 11.6.2. Occupational Health Since AR4, much has been written on the effects of heat on working people (Dunne et al., 2013; Kjellstrom et al., 2009a) and on other climate-related occupational health risks (Bennett and McMichael, 2010; Schulte and Chun, 2009). 11.6.2.1. Heat Strain and Heat Stroke Worldwide, more than half of all non-household labor-hours occur outdoors, mainly in agriculture and construction (International Fund for Agricultural Development, 2010; International Labor Organization, 2013). Individuals who are obliged to work outside in hot conditions, without access to shade, or sufficient water, are at heightened risk of heat-strain (ICD code T.67, “heat exhaustion”) and heat stroke. Health risks increase with the level of physical exertion. Agricultural and construction workers in tropical developing countries are therefore among the most exposed, but heat stress is also an issue for those working indoors in environments that are not temperature- controlled, and even for some workers in high-income countries such as the USA (Luginbuhl et al., 2008). (See Figure 11-5). Moreover, at higher temperatures there is potential conflict between health protection and economic productivity (Kjellstrom et al., 2011): as workers take longer rests to prevent heat stress, hourly productivity goes down (Sahu et al., 2013). [INSERT FIGURE 11-5 HERE Figure 11-5: The 1980-2009 average of the hottest months globally, measured in web bulb globe temperature (WBGT), which combines temperature, humidity, and other factors into a single index of the impact on work capacity and threat of heat exhaustion. The insert shows the International Standard Organization standard (1989) for heat stress in the workplace that leads to recommendations for increased rest time per hour to avoid heat exhaustion at different work levels. This is based on studies of healthy young workers and includes a margin of safety. Note that some parts of the world already exceed the level for safe work activity during the hottest month. In general, with climate change, for every 1°C that Tmax goes up, the WBGT goes up by about 0.9°C, leading to more parts of the world being restricted for more of the year, with consequent impacts on productivity, heat exhaustion, and need for air conditioning to protect health. Source: Lemke and Kjellstrom, 2012.] Subject to Final Copyedit 24 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 11.6.2.2. Heat Exhaustion and Work Capacity Loss There are international standards of maximum recommended workplace heat exposure and hourly rest time (e.g. (ISO, 1989; Parsons, 2003)) for both acclimatized and non-acclimatized people. In hot countries during the hot season, large proportions of the workforce are affected by heat, and the economic impacts of reduced work capacity may be sufficient to jeopardize livelihoods (Kjellstrom et al., 2011; Kjellstrom et al., 2009a; Kjellstrom and Crowe, 2011; Lecocq and Shalizi, 2007). Kjellstrom and Crow (2011) and Dunne et al. (Dunne et al., 2013) report that loss of work productivity during the hottest and wettest seasons has already occurred, at least in Asia and Africa. 11.6.2.3. Other Occupational Health Concerns In areas where vector-borne diseases, such as malaria and dengue fever, are common, people working in fields without effective protection may experience a higher incidence of these diseases when climatic conditions favour mosquito breeding and biting (Bennett and McMichael, 2010). Increasing heat exposure in farm fields during the middle of the day may lead to more work during dawn and dusk when some of the vectors are biting humans more actively. Exposure to heat affects psychomotor, perceptual, and cognitive performance (Hancock et al., 2007) and increases risk of injuries (Ramsey, 1995). Extreme weather events and climate-sensitive infectious diseases also pose occupational risks to health workers, which may in turn undermine health protection for the wider population (World Health Organization, 2009b). Other mechanisms include elevated occupational exposures to toxic chemical solvents which evaporate faster at higher temperatures (Bennett and McMichael, 2010) and rising temperatures reducing sea ice and increasing risk of drowning in those engaged in traditional hunting and fishing in the Arctic (Ford et al., 2008). 11.6.2.4. Near-Term Future Projections have been made of the future effects of heat on work capacity (Dunne et al., 2013; Kjellstrom et al., 2009b). Temperature and humidity were both included, and the modeling took into account the changes in the workforce distribution relating to the need for physical activity. In Southeast Asia, in 2050, the model indicates that more than half the afternoon work hours will be lost due to the need for rest breaks (Kjellstrom et al., 2013). By 2100, under RCP4.5, Dunne et al. (2013) project up to a 20% loss of productivity globally. There is an unfortunate trade-off between health impact and productivity, which creates risks for poor and disenfranchised laborers working under difficult working conditions and inflexible rules (Kjellstrom et al., 2011; Kjellstrom et al., 2009a; Sahu et al., 2013). 11.6.3. Mental Health Harsher weather conditions such as floods, droughts, and heat waves tend to increase the stress on all those who are already mentally ill, and may create sufficient stress for some who are not yet ill to become so (Berry et al., 2010). Manifestations of disaster-related psychiatric trauma include severe anxiety reactions (such as post-traumatic stress) and longer-term impacts such as generalised anxiety, depression, aggression, and complex psychopathology (Ahern et al., 2005; Ronan et al., 2008). For slow-developing events such as prolonged droughts, impacts include chronic psychological distress and increased incidence of suicide (Alston and Kent, 2008; Hanigan et al., 2012). Extreme weather conditions may have indirect effects on those with mental illness, through the impacts on agricultural productivity, fishing, forestry and other economic activities. Disasters such as cyclones, heat waves and major floods may also have destructive effects in cities. Here again, the mentally ill may be at risk: cities often feature zones of concentrated disadvantage where mental disorders are more common (Berry, 2007) and there is also higher risk of natural disasters (such as flooding). Subject to Final Copyedit 25 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 In addition to effects of extreme weather events on mental health via the risk/disadvantage cycle, there may be a distressing sense of loss, known as ‘solastalgia,’ that people experience when their land is damaged (Albrecht et al., 2007) and they lose amenity and opportunity. 11.6.4. Violence and Conflict Soil degradation, freshwater scarcity, population pressures and other forces that are related to climate are all potential causes of conflict. The relationships are not straightforward, however, as many factors influence conflict and violence. The topic is reviewed closely in Chapter 12, which concludes that factors associated with risk of violent conflict, such as poverty and impaired state institutions, are sensitive to climate variability, but evidence of an effect of climate change on violence is contested. Also, it is noted that populations affected by violence are particularly vulnerable to the impacts of climate change on health and social well-being. 11.7. Adaptation to Protect Health Climate change may threaten the progress that has been made in reducing the burden of climate-related disease and injury. The degree to which programs and measures will need modification to address additional pressures from climate change will depend on the current burden of ill-health, the effectiveness of current interventions, projections of where, when, and how the health burden could change with climate change, the feasibility of implementing additional programs, other stressors that could increase or decrease resilience, and the social, economic, and political context for intervention (Ebi et al., 2006). The scientific literature on adaptation to climate change has expanded since AR4, and there are many more national adaptation plans that include health, but investment in specific health protection activities is growing less rapidly. A review by the World Health Organization in 2012 estimated that commitments to health adaptation internationally amount to less than 1% of the annual health costs attributable to climate change in 2030 (World Health Organization Regional Office for Europe, 2013). The value of adaptation is demonstrated by the health impacts of recent disasters associated with extreme weather and climate events, although not necessarily attributed with confidence to climate change itself. For example, approximately 500,000 people died when cyclone Bhola (category 3 in severity) hit East Pakistan (present day Bangladesh) in 1970. In 1991, a cyclone of similar severity caused about 140,000 deaths. In November 2007, cyclone Sidr (category 4) resulted in approximately 3,400 deaths; the population had grown by more than 30 million in the intervening period (Mallick et al., 2005). Bangladesh achieved this remarkable reduction in mortality through effective collaborations between governmental and non-governmental organizations and local communities (Khan, 2008). Alongside improving general disaster education (greatly assisted by rising literacy rates, especially among women), the country deployed early warning systems and built a network of cyclone shelters. Early warning systems included high technology information systems and relatively simple measures such as training volunteers to distribute warning messages by bicycle. Efforts to adapt to the health impacts of climate change can be categorized as incremental, transitional, and transformational actions (O'Brien et al., 2012). Incremental adaptation includes improving public health and health care services for climate-related health outcomes, without necessarily considering the possible impacts of climate change. Transitional adaptation means shifts in attitudes and perceptions, leading to initiatives such as vulnerability mapping and improved surveillance systems that specifically integrate environmental factors. Transformational adaptation (see Chapter 16), which requires fundamental changes in systems, has yet to be implemented in the health sector. Subject to Final Copyedit 26 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 11.7.1. Improving Basic Public Health and Health Care Services Although the short time period since health adaptation options have been implemented means evidence of effectiveness in specifically reducing climate change-related impacts is currently lacking, there is abundant evidence of steps that may be taken to improve relevant public health functions (Woodward et al., 2011). This is important because the present health status of a population may be the single most important predictor of both the future health impacts of climate change and the costs of adaptation (Pandey, 2010). Most health adaptation focuses on improvements in public health functions to reduce the current adaptation deficit, such as enhancing disease surveillance, monitoring environmental exposures, improving disaster risk management, and facilitating coordination between health and other sectors to deal with shifts in the incidence and geographic range of diseases (Woodward et al., 2011). Examples of incremental health care interventions include introduction of vaccination programs in the United States, after which seasonal outbreaks of rotavirus, a common climate-sensitive pathogen, were delayed and diminished in magnitude (Tate et al., 2009). Post-disaster initiatives also are important. For example, an assessment of actions to improve the resilience of vulnerable populations to heat waves recommended staff planning over the summer period, cooling of health care facilities, training of staff to recognize and treat heat strain, and monitoring of those in the highest risk population groups (World Health Organization Regional Office for Europe, 2009). Ensuring essential medical supplies for care of individuals with chronic conditions, including effective post-disaster distribution, would increase the ability of communities to manage large-scale floods and storms. In Benin, one measure proposed as part of the national response to sea level rise and flooding is expanded health insurance arrangements, so that diseases such as malaria and enteric infections can be treated promptly and effectively (Dossou and Glehouenou-Dossou, 2007). 11.7.2. Health Adaptation Policies and Measures Transitional adaptation moves beyond focusing on reducing the current adaptation deficit to considerations of how a changing climate could alter health burdens and the effectiveness of interventions (Frumkin et al., 2008). For example, maintaining and improving food safety in the face of rising temperatures and rainfall extremes depends on effective interactions between human health and veterinary authorities, integrated monitoring of food-borne and animal diseases, and improved methods to detect pathogens and contaminants in food (Tirado et al., 2010). Indicators of community functioning and connectedness also are relevant because communities with high levels of social capital tend to be more successful in disseminating health and related messages, providing support to those in need (Frumkin et al., 2008). Vulnerability Mapping Vulnerability mapping is being increasingly used to better understand current and possible future risks related to climate change. For example, Reid et al. (2009) mapped community determinants of heat vulnerability in the U.SA. The four factors explaining most of the variance were a combination of social and environmental factors, social isolation, prevalence of air conditioning, and the proportion of the population who were elderly or diabetic. Remote sensing technologies are now sufficiently fine-grained to map local vulnerability. For example, these technologies can be used to map surface temperatures and urban heat island effects at the neighborhood scale, indicating where city greening and other urban cooling measures could be most effective, and alerting public health authorities to populations that may be at greatest risk of heat waves (Luber and McGeehin, 2008). In another example, spatial modeling of geo-referenced climate and environmental information was used to identify characteristics of domestic malaria transmission in 2009-2012 in Greece, to guide malaria control efforts (Sudre et al., 2013). Mapping at regional and larger scales may be useful to guide adaptation actions. In Portugal, modeling of Lyme disease indicates that future conditions will be less favorable for disease transmission in the south, but more favorable in the center and northern parts of the country (Casimiro et al., 2006). This information can be used to modify surveillance programs before disease outbreaks occur. To capture a more complete picture of vulnerability, mapping exercises Subject to Final Copyedit 27 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 also could consider climate sensitivity and adaptation capacity, such as was done in an assessment of climate change and risk of poverty in Africa (Thornton et al., 2008). 11.7.3. Early Warning Systems Early warning systems have been developed in many areas to prevent negative health impacts through alerting public health authorities and the general public about climate-related health risks. Effective early warning systems take into consideration the range of factors that can drive risk and are developed in collaboration with end users. Components of effective early warning systems include forecasting weather conditions associated with increased morbidity or mortality, predicting possible health outcomes, identifying triggers of effective and timely response plans that target vulnerable populations, communicating risks and prevention responses, and evaluating and revising the system to increase effectiveness in a changing climate (Lowe et al., 2011). Heat wave early warning systems are being increasingly implemented, primarily in high-income countries. Of eight studies of the effectiveness of heat wave early warning systems or heat prevention activities to reduce heat-related mortality, seven reported fewer deaths during heat waves after implementation of the system (Chau et al., 2009; Ebi et al., 2004; Fouillet et al., 2008; Palecki et al., 2001; Tan et al., 2007; Weisskopf et al., 2002) ; only Morabito et al. (2012) was inconclusive. For example, in the summer of 2006, France experienced high temperatures with about 2,000 excess deaths. This was over 4,000 fewer deaths than was anticipated on the basis of what occurred in the 2003 heat wave. A national assessment attributed the lower than expected death toll to greater public awareness of the health risks of heat, improved health care facilities, and the introduction in 2004 of a heat wave early warning system (Fouillet et al., 2008). A review of the heat wave early warning systems in the twelve European countries with such plans concluded that evaluations of the effectiveness of these systems is urgently needed to inform good practices, particularly understanding which actions increase resilience (Lowe et al., 2011). Early warning systems have been developed also for vector-borne and food-borne infections, although evidence of their effectiveness in reducing disease burdens is limited. In Botswana, an early warning system forecasts malaria incidence up to 4 months in advance based on observed rainfall; inter-annual and seasonal variations in climate are associated with outbreaks of malaria in this part of Africa. Model outputs include probability distributions of disease risk and measures of the uncertainty associated with the forecasts (Thomson et al., 2006). A weather-based forecasting model for dengue, developed in Singapore, predicted epidemics 13 months ahead of the peak in new cases, which gave the national control program time to increase control measures (Hii et al., 2012). A study of campylobacteriosis in the United States developed models of monthly disease risk with a very good fit in validation data sets (R2 up to 80%) (Weisent et al., 2010). 11.7.4. Role of Other Sectors in Health Adaptation Other sectors, including ecosystems, water supply and sanitation, agriculture, infrastructure, energy and transportation, land use management, and others, play an important part in determining the risks of disease and injury resulting from climate change. Within the context of the EuroHEAT project, a review of public health responses to extreme heat in Europe identified transport policies, building design, and urban land use as important elements of national and municipal heat wave and health action plans (World Health Organization Regional Office for Europe, 2009). A study examining well-established interventions to reduce the urban heat island effect (replacing bitumen and concrete with more heat-reflective surfaces, and introducing more green spaces to the city) estimated these would reduce heat- related emergency calls for medical assistance by almost 50% (Silva et al., 2010). Urban green spaces lower ambient temperatures, improve air quality, provide shade, and may be good for mental health (van den Berg et al., 2010). However, the extent to which changes in these factors reduce heat wave-related morbidity and mortality depend on location. A study in London, UK, found that built form and other dwelling characteristics more strongly influenced indoor temperatures during heat waves than did the urban health island effect (Oikonomou and Wilkinson, 2012). Subject to Final Copyedit 28 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 A review of food aid programs indicates that a rapid response to the risk of child under-nutrition, targeted to those in greatest need, with flexible financing and the capacity to rapidly scale-up depending on need, may reduce damaging health consequences (Alderman, 2010). Community-based programs designed for other purposes can facilitate adaptation, including disaster risk management. In the Philippines, for example, interventions in low-income urban settings with the potential to reduce the harmful effects of climate extremes on health include savings schemes, small-scale loans, hygiene education, local control and maintenance of water supplies, and neighborhood level solid waste management strategies (Dodman et al., 2010). It is important to note that climate change adaptation in other sectors may influence health in a positive manner (eg re-vegetation of watersheds to improve water quality), or on occasion, exacerbate health risks (eg urban wet-lands designed primarily for flood control may promote mosquito breeding) (Medlock and Vaux, 2011). 11.8. Adaptation Limits under High Levels of Warming Most attempts to quantify health burdens associated with future climate change consider modest increases in global average temperature, typically less than 2° C. However, research published since AR4 raises doubt over whether it will be possible to limit global warming to 2°C above pre-industrial temperatures (Anderson and Bows, 2011; PriceWaterhouseCoopers, 2012; Rogelj et al., 2009). It is therefore increasingly important to examine the likely health consequences of warming beyond 2°, including extreme warming of 4-6°C or higher. Predictions of this nature are limited by uncertainty about climatic as well as key, non-climatic determinants of health including the nature and degree of adaptation. Here, we instead focus primarily on physiological or ecological limits that constrain our ability to adapt and protect human health and wellbeing (Section 16.4.1). It can be assumed that the increase in many important climate-related health impacts at increasingly higher levels of warming will be greater than simple linear increments; that is, that the health consequences of a 4°C temperature increase will be more than twice those of a +2°C world (see Figure 11-6). Nonlinear and threshold effects have been observed in the mortality response to extreme heat (Anderson and Bell, 2011; McMichael, 2013a), agricultural crop yields, as key determinants of childhood nutrition and development (Lobell et al., 2011a; Schlenker and Roberts, 2009), and infectious diseases (Altizer et al., 2006) , for example. These are also briefly elaborated here. [INSERT FIGURE 11-6 HERE Figure 11-6: Conceptual presentation of the health impacts from climate change and the potential for impact reduction through adaptation. Impacts are identified in eight health-related sectors based on assessment of the literature and expert judgments by authors of Chapter 11.The width of the slices indicates in a qualitative way the relative importance in terms of burden of ill-health globally at present and should not be considered completely independent. Impact levels are presented for the near-term era of committed climate change (2030-2040), in which projected levels of global mean temperature increases do not diverge substantially across emissions scenarios. For some sectors, e.g., vector-borne diseases, heat/cold stress, and agricultural production and undernutrition, there may be benefits to health in some areas, but the net impact is expected to be negative. Estimated impacts are also presented for the longer-term era of climate options (2080-2100), for global mean temperature increase of 4°C above preindustrial levels, which could potentially be avoided by vigorous mitigation efforts taken soon. For each timeframe, impact levels are estimated for the current state of adaptation and for a hypothetical highly adapted state, indicated by different colors.] 11.8.1. Physiological Limits to Human Heat Tolerance In standard (or typical) conditions, core body temperatures will reach lethal levels under sustained periods of wet- bulb temperatures above about 35°C (Sherwood and Huber, 2010). Sherwood and Huber (2010) conclude that a global mean warming of roughly 7°C above current temperatures would create small land areas where metabolic heat dissipation would become impossible. An increase of 11-12°C would enlarge these zones to encompass most of the areas occupied by today’s human population. Subject to Final Copyedit 29 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 The above analysis is likely a conservative estimate of an absolute limit to human heat tolerance because working conditions are hazardous at lower thresholds. The U.S. military, for example, suspends all physical training and strenuous exercise when the wet bulb globe temperature (WBGT) exceeds 32°C (Willett and Sherwood, 2012) while international labor standards suggest the time acclimatized individuals spend doing low intensity labor such as office work be halved under such conditions (Kjellstrom et al., 2009a).1 One estimate suggests global labor productivity will be reduced during the hottest months to 60% in 2100 and less than 40% in 2200 under the RCP 8.5 scenario in which global mean temperatures rise 3.4°C by 2100 and 6.2°C by 2200 relative to 1861-1960 (Dunne et al., 2013). It is projected that tropical and mid-latitude regions including India, Northern Australia, Southeastern USA will be particularly badly affected (Dunne et al., 2013; Willett and Sherwood, 2012). [FOOTNOTE 1: WBGT is a heat index closely related to the wet-bulb temperature that also incorporates measures of radiant heat from the Sun and evaporative cooling due to wind.] 11.8.2. Limits to Food Production and Human Nutrition Agricultural crops and livestock similarly have physiological limitations in terms of thermal and water stress. For example, production of the staple crops maize, rice, wheat and soybean is generally assumed to face an absolute temperature limit in the range of 40-45°C (Teixeira et al., 2011), while key phenological stages such as sowing to emergence, grain-filling, and seed set have maximum temperature thresholds near or below 35°C (Porter and Semenov, 2005; Porter and Gawith, 1999; Yoshida et al., 1981). The existence of critical climatic thresholds and evidence of non-linear responses of staple crop yields to temperature and rainfall (Brázdil et al., 2009; Lobell et al., 2011b; Schlenker and Roberts, 2009) thus suggest that there may be a threshold of global warming beyond which current agricultural practices can no longer support large human civilizations, and the impacts on malnourishment and undernutrition described in Section 11.6.1 will become much more severe. However, current models to estimate the human health consequences of climate-impaired food yields at higher global temperatures generally incorporate neither critical thresholds nor nonlinear response functions (Lake et al., 2012; Lloyd et al., 2011), reflecting uncertainties about exposure-response relations, future extreme events, the scale and feasibility of adaptation, and climatic thresholds for other influences such as infestations and plant diseases. Extrapolation from current models nevertheless suggests that the global risk to food security becomes very severe under an increase of 4-6°C or higher in global mean temperature (medium evidence, high agreement) (Chapter 7, Executive Summary). 11.8.3. Thermal Tolerance of Disease Vectors Substantial warming in higher-latitude regions will open up new terrain for some infectious diseases that are limited at present by low temperature boundaries, as already evidenced by the northward extensions in Canada and Scandinavia of tick populations, the vectors for Lyme disease and tick-borne encephalitis (Lindgren and Gustafson, 2001; Ogden et al., 2006). On the other hand, the emergence of new temperature regimes that exceed optimal conditions for vector and host species will reduce the potential for infectious disease transmission and, with high enough temperature rise, may eventually eliminate some infectious diseases that exist at present close to their upper tolerable temperature limits. For example, adults of two malaria-transmitting mosquito species are unable to survive temperatures much above 40°C in laboratory experiments (Lyons et al., 2012), although in the external world they may seek out tolerable microclimates. Reproduction of the malaria parasite within the mosquito is impaired at lesser raised temperatures (Paaijmans et al., 2009). Larval development of Aedes albopictus, an Asian mosquito vector of dengue and chikungunya, also does not occur at or above 40°C (Delatte et al., 2009). 11.8.4. Displacement & Migration under Extreme Warming Weather extremes and longer term environmental change including sea level rise lead to both more people displaced and increase in populations that are effectively trapped (Section 12.4.1.2). This trend is expected to be more pronounced under extreme levels of warming (Section 16.5). Gemenne (2011) argues that the most significant difference between the nature of human migration in response to 4°C of warming relative to 2°C would be to Subject to Final Copyedit 30 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 remove many people’s ability to choose whether to stay or leave when confronted with environmental changes. Health studies of refugees, migrants, and people in resettlement schemes suggest that forced displacement, in turn, is likely to lead to more adverse health impacts than voluntary migration or planned resettlement (McMichael et al., 2012). The health risks associated with forced displacement include undernutrition; food- and water-borne illnesses; diseases related to overcrowding such as measles, meningitis and acute respiratory infections; sexually-transmitted diseases; increased maternal mortality; and mental health disorders (McMichael et al., 2012). 11.8.5. Reliance on Infrastructure Under severe climate regimes, societies may be able to protect themselves by enclosing places for living and working, first for their most vulnerable members: the young, old, ill, and manual laborers. This strategy will mean increased vulnerability to infrastructure failure and unreliable energy and water supplies. Electrical power outages have been linked to both accidental and disease-related deaths in temperate climates (Anderson and Bell, 2012), and failures in power supplies are more likely to occur during extreme weather events (Section 19.6.2.1). Large-scale reliance on air-conditioning under a significantly hotter climate regime would therefore pose a serious health risk. 11.9. Co-Benefits Essentially every human activity affects (and is affected by) climate and health status in some way, but not all are strongly linked to either and even fewer strongly to both. Here we focus on measures to mitigate the atmospheric concentration of warming climate altering pollutants (CAPs) that also hold the potential to significantly benefit human health. These so-called co-benefits include health gains from strategies that are directed primarily at climate change, and mitigation of climate change from well-chosen policies for health advancement (Apsimon et al., 2009; Haines et al., 2007; Shindell et al., 2012; Smith and Balakrishnan, 2009; UNEP, 2011). The literature on health co- benefits associated with climate change mitigation strategies falls into several categories (Smith and Balakrishnan, 2009; Smith et al., 2009). These include: Reduce emissions of health-damaging pollutants, either primary or precursors to other pollutants in association with changes in energy production, energy efficiency, or control of landfills; Increase access to reproductive health services; Decrease meat consumption (especially from ruminants) and substitution of low-carbon healthy alternatives; Increase active transport particularly in urban areas; Increase urban green-space. In addition, although not discussed here, there are potential health side effects of mitigation measures, such as geoengineering, biofuel expansion, and carbon taxes that are potentially deleterious for human health (Tilman et al., 2009). See WGII Ch 19. In Table 11-3, we summarize what is known about the main categories of co-benefits, but because of space limitation, we only provide additional detail for two of them below. [INSERT FIGURE 11-7 Figure 11-7: Illustrative co-benefits comparison of the health and climate cost-effectiveness of selected household, transport, and power sector interventions (Smith and Haigler, 2008). Area of each circle denotes the total social benefit in international dollars from the combined value of carbon offsets (valued at 10$/tCO2e) and averted DALYs [$7450/DALY, which is representative of valuing each DALY at the average world GDP (PPP) per capita in 2000]. The vertical bar shows the range of the cut offs for cost-effective and very cost-effective health interventions in India and China using the WHO CHOICE criteria (World Health Organization, 2003). This figure evaluates only a small subset of all co-benefits opportunities and thus should not be considered either current or complete. It does illustrate, however, the kind of comparisons that can help distinguish and prioritize options. Note that even with the log-log scaling, there are big differences among them. For other figures comparing the climate and health benefits of co-benefits actions including those in food supply and urban design, see Haines et al. (Haines et al., 2009). See the original reference for details of the calculations in this figure (Smith and Haigler, 2008).] [INSERT TABLE 11-3 HERE Table 11-3: Examples of recent (post AR4) research studies on co-benefits of climate change mitigation and public health policies. For recent estimates of the global and regional burden of disease from the various risk factors involved, see Lim et al. (2012).] Subject to Final Copyedit 31 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 11.9.1. Reduction of Co-Pollutants Most of the publications related to CAPs and health-damaging pollutants refer to fuel combustion and fall into three major categories: 1) improvement in energy efficiency will reduce emissions of CO2 and health-damaging pollutants, providing these gains are not outpaced by increases in energy demand, and the energy is derived from combustion of fossil fuels or non-renewable biomass fuels, either directly or through the electric power system; 2) increases of combustion efficiency (decreasing emission of incomplete combustion products) will have both climate and health benefits, even if there is no change in energy efficiency and/or fuel itself is renewable, because a number of the products of incomplete combustion are climate altering and nearly all are damaging to health (Smith and Balakrishnan, 2009); and 3) Increased use of non-combustion sources, such as wind, solar, tidal, wave and geothermal energy, would reduce emissions of warming CAPs and health damaging air pollutants, providing benefits for climate and health (Jacobson et al., 2013). Studies of the health co-benefits of reduction in air pollutants include sources that produce outdoor air pollution (Bell et al., 2008) and household sources (Po et al., 2011). In many parts of the world, household fuel (poorly combusted biomass and coal) is responsible for a substantial percent of primary outdoor fine particle pollution as well, perhaps a quarter in India, for example (Lim et al., 2012). In many parts of the world, household fuel (poorly combusted biomass and coal) is responsible for much fine particle outdoor air pollution and may contribute to long- range transport of hazardous air pollutants (Anenberg et al., 2013). This indicates that reductions in emissions from household sources will yield co-benefits through the outdoor pollution pathway as well. If interventions result in reductions in coal combustion, there are a range of other potential health benefits beyond reduction of particulate air pollution emissions, including reducing other types of health-damaging emissions and the human impacts from coal mining (Lockwood, 2012; Smith et al., 2013). Another category of air pollution co-benefits comes from controls on methane emissions that both reduce radiative forcing and potentially reduce human exposures to ambient ozone, for which methane is a precursor. 11.9.1.1. Outdoor Sources Primary co-pollutants, such as particulate matter (PM) and carbon monoxide (CO) are those released at the point of combustion, while secondary co-pollutants, such as tropospheric ozone and sulfate particles, are formed downwind from the combustion source via atmospheric chemical interactions (Jerrett et al., 2009) and can be transported long distances. The burden of disease from outdoor exposures in a country may often be greater in populations with low socioeconomic status, both because of living in areas with higher exposures and because these populations often have worse health and are subjected to multiple additional negative environmental and social exposures (Morello- Frosch et al., 2011). 11.9.1.2. Household Sources Globally, the largest exposures from the pollutants from poor fuel combustion occur in the poorest populations. This is because household use of biomass for cooking is distributed nearly inversely with income. Essentially, no poor family can afford gas or electricity for cooking and very few families who can afford to do so, do not. Thus, the approximate 41% of all world households using solid fuels for cooking are all among the poor in developing countries (Bonjour et al., 2013). Although biomass makes up the bulk of this fuel and creates substantial health impacts from products of incomplete combustion when burned in simple stoves (Lim et al., 2012), probably the greatest health and largest climate impacts per household result from use of coal, which can also be contaminated with sulfur and a range of toxic elements as well (Edwards et al., 2004; Zhang and Smith, 2007). Successfully accelerating the reduction of impacts from these fuels, however, has not been found to be easily accomplished with Subject to Final Copyedit 32 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 biomass/coal stove programs implemented to date and may require moving to clean fuels (Bruce et al., 2013). The climate benefits from improving household biomass fuel combustion come in part from potential reduction of net warming by reducing emissions of aerosols (including black carbon), but more confidently from reduction of CH4 and other CAPs that are produced by incomplete combustion, as well as reductions in net CO2 emissions if interventions are applied in areas relying on non-renewably harvested wood fuel (WG1, Section 8.5.3). 11.9.1.3. Primary Co-Pollutants Outdoor exposure to PM, especially to particles with diameters less than 2.5 µm (PM2.5), contributes significantly to ill-health including cardio- and cerebrovascular disease, adult chronic and child acute respiratory illnesses, lung cancer, and possibly other diseases. The Comparative Risk Assessment (CRA) for outdoor air pollution done as part of the Global Burden of Disease (GBD) 2010 Project found approximately 3.2 million premature deaths globally from ambient particle pollution or about 3% of the global burden of disease (Lim et al., 2012). Importantly, reductions in ambient PM concentrations have also been shown to decrease morbidity and premature mortality (Boldo et al., 2010). A significant portion of ambient particle pollution derives from fuel combustion, perhaps 80% globally (GEA, 2012). Because of higher exposures, an additional set of diseases has also been associated with combustion products in households burning biomass and/or coal for cooking and heating. Thus, in addition to the diseases noted above, cataracts, low birth weight, and stillbirth have been associated strongly with exposures to incomplete combustion products, such as PM and CO. CO has impacts on unborn children in utero through exposures to their pregnant mothers (World Health Organization Regional Office for Europe, 2010). There is also growing evidence of exacerbation of tuberculosis (Pokhrel et al., 2010) in adults and cognitive effects in children (Dix-Cooper et al., 2012). The CRA of the GBD-2010 found 3.5 million premature deaths annually from household air pollution derived from cooking fuels or 4.4% of the global burden of disease (Lim et al., 2012). Importantly, there are also studies showing health benefits of household interventions, for child pneumonia (Smith et al., 2011) blood pressure (Baumgartner et al., 2011; McCracken et al., 2007) , lung cancer (Lan et al., 2002) , and chronic obstructive pulmonary disease (Chapman et al., 2005). Another half a million premature deaths are attributed to household cookfuel’s contribution to outdoor air pollution, making a total of about 4 million in 2010 or 4.9% of the global burden of disease (Lim et al., 2012). Black carbon (BC), a primary product of incomplete combustion, is both a strong CAP and health-damaging (Bond et al., 2013; IPCC, 2007; Ramanathan and Carmichael, 2008). A systematic review, meta-analysis, and the largest cohort study to date of the health effects of BC found that there were probably stronger effects on mortality from exposure to BC than for undifferentiated fine particles (PM2.5) (Smith et al., 2009). Reviews have concludes that abatement of particle emissions including BC represents an opportunity to achieve both climate mitigation and health benefits (Shindell et al., 2012; UNEP, 2011). WG1 (Box TS-6), however, concluded that the net impact of BC emissions reductions overall is not certain as to sign, i.e., whether net warming or cooling. Nevertheless, there would be climate (and health) benefits in circumstances where BC is emitted without many other cooling aerosols, as with diesel and kerosene combustion (Lam et al., 2012). Other examples of climate forcing, health-damaging co-pollutants of CO2 from fuel use are carbon monoxide, non- methane hydrocarbons, and sulfur and nitrogen oxides. Each co-pollutant poses risks as well as being climate altering in different ways. See WGI for more on climate potential and WHO reviews of health impacts (World Health Organization Regional Office for Europe, 2010; World Health Organization, 2006). 11.9.1.4. Secondary Co-Pollutants In addition to being a strong GHG, methane (CH4) is also a significant precursor to regional anthropogenic tropospheric ozone production, which itself is both a GHG and damaging to health, crops, and ecosystems (WG1 TS.5.4.8). Thus, reductions in CH4 could lead to reductions in ambient tropospheric ozone concentrations, which in turn could result in reductions in population morbidity and premature mortality and climate forcing. Subject to Final Copyedit 33 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 One study found that a reduction of global anthropogenic CH4 emissions by 20% beginning in 2010 could decrease the average daily maximum 8-h surface ozone by 1 ppb by volume, globally; sufficient to prevent 30,000 premature all-cause mortalities globally in 2030, and 370,000 between 2010 and 2030 (West, Fiore et al. 2012). CH4 emissions are generally accepted as the primary anthropogenic source of tropospheric ozone concentrations above other human-caused emissions of ozone precursors (West et al., 2007) and thus, the indirect health co-benefits of CH4 reductions are epidemiologically significant. On the other hand, work done for the GBD-2010 estimated 150,000 premature deaths from all ozone exposures globally in 2010, indicating a more conservative interpretation of the evidence for mortality from ozone (Lim et al., 2012). In an analysis of ozone trends from 1998-2008 in the United States, Lefohn et al. (2010) found that 1-hour and 8- hour ambient ozone averages have either decreased or failed to increase due to successful regulations of ozone precursors, predominantly NOx and CH4. This is consistent with the US EPA (2010) conclusion that in the US, for the period 1980-2008, emissions of nitrogen oxides and volatile organic compounds fell by 40% and 47%, respectively (Lefohn et al., 2010; US EPA, 2010). These results point to the effectiveness of reducing ambient ozone concentrations through regulatory tools that reduce the emissions of ozone precursors, some of which, like CH4, are GHGs. Not every CAP emitted from fuel combustion is warming. The most prominent example is sulfur dioxide emitted from fossil fuel combustion, which changes to particle sulfate in the atmosphere. Although health damaging, sulfate particles have a cooling effect on global radiative forcing. Thus, reduction of sulfur emissions, which is important for health protection, does not qualify as a co-benefit activity since it actually acts to unmask more of the warming effect of other CAP emissions (Smith et al., 2009). 11.9.1.5. Case Studies of Co-Benefits of Air Pollution Reductions A recent UNEP- and WMO-led study of black carbon and tropospheric ozone found that, if all of 400 proposed BC and CH4 mitigation measures were implemented on a global scale, the estimated benefits to health would come predominately from reducing PM2.5 (0.7 – 4.6 million avoided premature deaths; 5.3 – 37.4 million avoided years of life lost) compared to tropospheric ozone (0.04 – 0.52 million avoided premature deaths; 0.35 – 4.7 million avoided years of life lost) based on 2030 population figures (UNEP, 2011). About 98% of the avoided deaths would come from reducing PM2.5, with 80% of the estimated health benefits occurring in Asia (Anenberg et al., 2012). Another study of the reduction of PM and ozone exposures due to CAPs emissions controls and including climate change feedback showed potential reductions of 1.3 million premature deaths by 2050 with avoided costs of premature mortality many times those of the estimated cost of abatement (West et al., 2013). A study of the benefits of a hypothetical 10-year program to introduce advanced combustion cookstoves in India found that in addition to reducing premature mortality by about 2 million and DALYs by 55 million over that period, there would be reduction of 0.5-1.0 billion tons CO2-eq (Wilkinson et al., 2009). Another study of India found a potential to reduce 570 thousand premature deaths a year, one-third of national BC emissions, and 4% of all national greenhouse emissions by hypothetical substitution of clean household fuel technologies (Venkataraman et al., 2010). In their estimation of effects of hypothetical physical and behavioral modifications in UK housing, Wilkinson and colleagues (Wilkinson et al., 2009) found that the magnitude and direction of implications for health depended heavily on the details of the intervention. However, the interventions were found to be generally positive for health. In a strategy of housing modification that included insulation, ventilation control, and fuel switching, along with behavioral changes, it was estimated that 850 fewer DALYs, and a savings of 0.6 megatonnes of CO2 per million population in one year could be achieved. These calculations were made by comparing the health of the 2010 population with and without the specified physical and behavioral modifications (Wilkinson et al., 2009). Markandya et al. (2009) assessed the changes in emissions of PM2.5 and subsequent effects on population health that could result from climate change mitigation measures aimed to reduce GHG emissions by 50% by 2050 (compared Subject to Final Copyedit 34 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 with 1990 emissions) from the electricity generation sector in the EU, China, and India In all three regions, changes in modes of production of electricity to reduce CO2 emissions were found to reduce PM2.5 and associated mortality. The greatest effect was found in India and the smallest in the EU. The analysis also found that if the health benefits were valued similarly to the approach used by the EU for air pollution, they offset the cost of GHG emission reductions, especially in the Indian context where emissions are high but costs of implementing the measures are low (Markandya et al., 2009). 11.9.2. Access to Reproductive Health Services Population growth influences the consumption of resources and emissions of CAPs (Cohen, 2010). Although population growth rates and total population size do not alone determine emissions, population size is an important factor. One study showed that CO2 emissions could be lower by 30% by 2100 if access to contraception was provided to those women expressing a need for it (O'Neill et al., 2010). Providing the unmet need for these services in areas such as the Sahel region of Africa that has both high fertility and high vulnerability to climate change can potentially significantly reduce human suffering as climate change proceeds (Potts and Henderson, 2012). This is important not only in poor countries, however, but also some rich ones like the US, where there is unmet need for reproductive health services as well as high CO2 emissions per capita (Cohen, 2010). Also, because of income rise in developing countries and concurrent reduction of greenhouse emissions in developed countries, a convergence in emissions per capita is expected in most scenarios by 2100 (WG1 TS5.2). Slowing population growth through lowering fertility, as might be achieved by increasing access to family planning, has been associated with improved maternal and child health – the co-benefit - in two main ways: increased birth spacing and reducing births by very young and old mothers. 11.9.2.1. Birth and Pregnancy Intervals Current evidence supports, with medium confidence, that short birth intervals (defined as birth intervals ≤24 months and inter-pregnancy intervals <6 months) are associated with increased risks of uterine rupture and bleeding (placental abruption and placenta previa) (Bujold et al., 2002; Conde-Agudelo et al., 2007). There is also a correlation between short birth interval and elevated risk of low-birth-weight (Zhu, 2005). Zhu (2005) found, in a review of three studies performed in the United States that the smallest risk of low birth weight was found with inter-pregnancy spacing between 18-23 months. Another review of five cohort studies found that a birth interval shorter than 18 months was significantly associated with decreased low birth weight, preterm birth, and infant mortality after controlling for confounding factors (Kozuki et al., 2013). Although an ecological analysis, a review across 17 countries shows a strikingly coherent picture of the relationship between birth spacing and reductions in child, infant and neonatal mortality (Figure 11-S3) with risk of child undernutrition and mortality both increasing with shorter birth intervals (Rutstein, 2005). One study estimated that shifting birth spacing from current patterns in the world to a minimum of 24 months would reduce by 20% (~2 million) the current excess child mortality in the world (Gribble et al., 2009; Rutstein, 2005). 11.9.2.2. Maternal Age at Birth Risk of death during delivery is highest in very young and very old mothers, and these are also the age groups that most often want to control their fertility (Engelman, 2010). Women who begin child bearing under the age of 20 years are at an increased risk of developing pregnancy complications such as cephalopelvic disproportion, obstructed labor, preterm delivery, toxemia, bleeding, and maternal death (Tsui et al., 2007). Additionally, children born to women under the age of 20 are at increased risk of fetal growth retardation and low birth weight, both of which can lead to long term physical and mental developmental problems (Tsui et al., 2007). Childbearing at later ages (>35 years) is associated with increased risk of miscarriage and other adverse health outcomes (Cleary- Goldman et al., 2005; Ujah et al., 2005). Subject to Final Copyedit 35 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Providing access to family planning saves women’s lives by reducing the total number of births and, in particular, through the reduction of births in high-risk groups (Prata, 2009) while simultaneously reducing total fertility and subsequent CAP emissions. Studies have found that when women have access to family planning, it is the highest risk age groups (youngest and oldest women) who reduce their fertility the most. In other words, family planning has a differential impact on maternal mortality reduction through reducing births in the highest risk groups (Diamond-Smith and Potts, 2011). 11.10. Key Uncertainties and Knowledge Gaps There is evidence that poverty alleviation, public health interventions such as provision of water and sanitation, and early warning and response systems for disasters and epidemics will help to protect health from climate risks. The key uncertainty is the extent to which society will strengthen these services, including taking into account the risks posed by climate change. With a strong response, climate change health effects are expected to be relatively small in the next few decades, but otherwise climate-attributable cases of disease and injury will steadily increase. Since AR4, national governments, through the World Health Assembly, have specifically called for increased research on (i) the scale and nature of health risks from climate change; (ii) effectiveness of interventions to protect health; (iii) health implications of adaptation and mitigation decisions taken in other sectors, (iv) improvement in decision support systems and surveillance, and (v) estimation of resource requirements. A recent scoping review identified quantitative peer-reviewed studies across all of these areas, with the exception of studies on the effectiveness or cost-effectiveness of targeted adaptation measures (Hosking and Campbell-Lendrum, 2012). There are also comparatively few studies of vulnerability in low and middle income populations, or of more complex disease pathways, such as the effect of more extreme weather on water and sanitation provision and diarrhea rates, on zoonotic diseases, or mental health. Studies of health co-benefits of climate change mitigation policies also remain rare compared to the size of the potential health gains. Potential negative side effects also need to be addressed, for example those arising from biofuel policies that compete with food production. Relevant research for health protection in the near term is therefore likely to come from cross-disciplinary studies, including public health decision makers, in the following areas; improved vulnerability and adaptation assessments that focus on particularly vulnerable populations and encompass complex causal pathways; quantitative estimation of the effectiveness of health adaptation measures; surveillance, monitoring, and observational systems that link climate, health and economic impact data and provide a basis for early warning systems as well as development of future scenarios; assessment of the health co-benefits of alternative climate mitigation policies. In the longer term, research will need to make the best use of traditional epidemiologic methods, while also taking into account the specific characteristics of climate change. These include the long-term and uncertain nature of the exposure and effects on multiple physical and biotic systems, with the potential for diverse and widespread effects, including high-impact events. There are low-probability, but plausible, scenarios for extreme climate regimes before the end of the century. Although difficult, it is important to develop robust methods to investigate the health implications of conditions that may apply in 2100, as decisions today about mitigation will determine their likelihood. Given the increase globally in life expectancies, many babies born this decade will be alive at the end of the century, and will be personally affected by the climate that is in place in 2100. Frequently Asked Questions FAQ 11.1: How does climate change affect human health? [to remain at the end of the chapter] Climate change affects health in three ways; 1) Directly, such as the mortality and morbidity (including “heat exhaustion”) due to extreme heat events, floods, and other extreme weather events in which climate change may play a role; 2) Indirect impacts from environmental and ecosystem changes, such as shifts in patterns of disease- carrying mosquitoes and ticks, or increases in waterborne diseases due to warmer conditions and increased precipitation and runoff; and 3) indirect impacts mediated through societal systems, such as undernutrition and Subject to Final Copyedit 36 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 mental illness from altered agricultural production and food insecurity, stress and undernutrition and violent conflict caused by population displacement, economic losses due to widespread “heat exhaustion” impacts on the workforce, or other environmental stressors, and damage to health care systems by extreme weather events. FAQ 11.2: Will climate change have benefits for health? [to remain at the end of the chapter] Yes. For example some populations in temperate areas may be at less risk from extreme cold, and may benefit from greater agricultural productivity, at least for moderate degrees of climate change. Some areas currently prone to flooding may become less so. However, the overall impact for nearly all populations and for the world as a whole is expected to be more negative than positive, increasingly so as climate change progresses. In addition, the latitude range in the world that may benefit from less cold (e.g. the far north of the Northern Hemisphere) has fewer inhabitants compared with the equatorial latitudes where the burden will be greatest. FAQ 11.3: Who is most affected by climate change? [to remain at the end of the chapter] While the direct health effects of extreme weather events receive great attention, climate change mainly harms human health by exacerbating existing disease burdens and negative impacts on daily life among those with the weakest health protection systems, and with least capacity to adapt. Thus, most assessments indicate that poor and disenfranchised groups will bear the most risk and, globally, the greatest burden will fall on poor countries, particularly on poor children, who are most affected today by such climate-related diseases as malaria, undernutrition, and diarrhea. However, the diverse and global effects of climate change mean that higher income populations may also be affected by extreme events, emerging risks, and the spread of impacts from more vulnerable populations. FAQ 11.4: What is the most important adaptation strategy to reduce the health impacts of climate change? [to remain at the end of the chapter] In the immediate future, accelerating public health and medical interventions to reduce the present burden of disease, particularly diseases in poor countries related to climatic conditions, is the single most important step that can be taken to reduce the health impacts of climate change. Priority interventions include improved management of the environmental determinants of health (such as provision of water and sanitation), infectious disease surveillance, and strengthening the resilience of health systems to extreme weather events. Alleviation of poverty is also a necessary condition for successful adaptation. There are limits to health adaptation, however. For example, the higher-end projections of warming indicate that before the end of the 21st Century, parts of the world would experience temperatures that exceed physiological limits during periods of the year, making it impossible to work or carry out other physical activity outside. FAQ 11.5: What are health “co-benefits” of climate change mitigation measures? [to remain at the end of the chapter] Many mitigation measures that reduce emissions of climate-altering pollutants (CAPs) have important direct health benefits in addition to reducing the risk of climate change. This relationship is called “co-benefits.” For example, increasing combustion efficiency in households cooking with biomass or coal could have climate benefits by reducing CAPs and at the same time bring major health benefits among poor populations. Energy efficiency and reducing reliance on coal for electricity generation not only reduces emissions of greenhouse gases, but also reduces emissions of fine particles which cause many premature deaths worldwide as well as reducing other health impacts from the coal fuel cycle. Programs that encourage “active transport” (walking and cycling) in place of travel by motor vehicle reduce both CAP emissions and offer direct health benefits. A major share of greenhouse gas emissions from the food and agriculture sector arises from cows, goats and sheep – ruminants that create the greenhouse gas methane as part of their digestive process. Reducing consumption of meat and dairy products from these animals may reduce ischemic heart disease (assuming replacement with plant-based polyunsaturates) and some types of cancer. Programs to provide access to reproductive health services for all women will not only lead to slower population growth and its associated energy demands, but also will reduce the numbers of child and maternal deaths. Subject to Final Copyedit 37 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 References Abuaku, B.K., J. Zhou, X. Li, S. Li, A. Liu, T. Yang, and H. Tan, 2009: Morbidity and mortality among populations suffering floods in Hunan, China: The role of socioeconomic status. Journal of Flood Risk Management, 2(3), 222-228. 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. Ahern, M.J., R.S. Kovats, P. Wilkinson, R. Few, and F. Matthies, 2005: Global health impacts of floods: epidemiological evidence. Epidemiol Rev, 27, 36-45. Albrecht, G., G.M. Sartore, L. Connor, N. Higginbotham, S. Freeman, B. Kelly, H. Stain, A. Tonna, and G. Pollard, 2007: Solastalgia: the distress caused by environmental change. Australas Psychiatry, 15 Suppl 1, S95-8. Alderman, H., 2010: Safety Nets Can Help Address the Risks to Nutrition from Increasing Climate Variability. The Journal of Nutrition, 140(1), 148S-152S. Alexander, K.A., M. Carzolio, D. Goodin, and E. Vance, 2013: Climate change is likely to worsen the public health threat of diarrheal disease in Botswana. International Journal of Environmental Research & Public Health, 10(4), 1202-30. Alonso, D., M.J. Bouma, and M. Pascual, 2011: Epidemic malaria and warmer temperatures in recent decades in an East African highland. Proc Biol Sci, 278(1712), 1661-9. Alston, M. and J. Kent, 2008: The Big Dry: The link between rural masculinities and poor health outcomes for farming men. Journal of Sociology, 44(2), 133-147. Altizer, S., A. Dobson, P. Hosseini, P. Hudson, M. Pascual, and P. Rohani, 2006: Seasonality and the dynamics of infectious diseases. Ecology Letters, 9(4), 467-484. Analitis, A., I. Georgiadis, and K. Katsouyanni, 2012: Forest fires are associated with elevated mortality in a dense urban setting. Occupational Environmental Medicine, 69(3), 158-162. Anderson, G.B. and M.L. Bell, 2011: Heat waves in the United States: mortality risk during heat waves and effect modification by heat wave characteristics in 43 U.S. communities. Environmental Health Perspectives, 119(2), 210-8. Anderson, G.B. and M.L. Bell, 2012: Lights out: impact of the August 2003 power outage on mortality in New York, NY. Epidemiology, 23(2), 189-93. Anderson, K. and A. Bows, 2011: Beyond ‘dangerous’ climate change: emission scenarios for a new world. Philosl Trans Roy Soc London - Series A, 369(1934), 20-44. Anderson, M., C. Carmichael, V. Murray, A. Dengel, and M. Swainson, 2012: Defining indoor heat thresholds for health in the UK. Perspectives in Public Health, 133(3), 158-164. Andreassen, A., S. Jore, P. Cuber, S. Dudman, T. Tengs, K. Isaksen, H. Hygen, H. Viljugrein, G. Anestad, P. Ottesen, and K. Vainio, 2012: Prevalence of tick borne encephalitis virus in tick nymphs in relation to climatic factors on the southern coast of Norway. Parasit Vectors, 5(1), 177. Andriopoulos, P., A. Economopoulou, and G. Spanakos, 2013: A local outbreak of autochthonous Plasmodium vivax malaria in Laconia, Greece—a re-emerging infection in the southern borders of Europe? International Journal of Infectious Disease, 17(2), e125-e128. Anenberg, S.C., J. Schwartz, D. Shindell, M. Amann, G. Faluvegi, Z. Klimont, G. Janssens-Maenhout, L. Pozzoli, R. van Dingenen, E. Vignati, L. Emberson, N.Z. Muller, J.J. West, M. Williams, V. Demkine, W.K. Hicks, J. Kuylenstierna, F. Raes, and V. Ramanathan, 2012: Global Air Quality and Health Co-Benefits of Mitigating Near-Term Climate Change through Methane and Black Carbon Emission Controls. Environ Health Perspect, 120(6), 831-839. Anenberg, S.C., K. Balakrishnan, J. Jetter, O. Masera, S. Mehta, J. Moss, and V. Ramanathan, 2013: Cleaner cooking solutions to achieve health, climate, and economic cobenefits. Environmental Science & Technology, 47(9), 3944-52. Angelini, P., P. Macini, A.C. Finarelli, C. Pol, C. Venturelli, R. Bellini, and M. Dottori, 2008: Chikungunya epidemic outbreak in Emilia-Romagna (Italy) during summer 2007. Parassitologia, 50(1-2), 97-98. Anyamba, A., K.J. Linthicum, and J.L. Small, 2012: Climate Teleconnections and Recent Patterns of Human and Animal Disease Outbreaks. PLoS Neglected Tropical Diseases, 6(1), e1465. Apsimon, H., M. Amann, S. Astrom, and T. Oxley, 2009: Synergies in addressing air quality and climate change. Climate Policy, 9(6), 669-680. Subject to Final Copyedit 38 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ari, T.B., A. Gershunov, R. Tristan, B. Cazelles, K. Gage, and N.C. Stenseth, 2010: Interannual Variability of Human Plague Occurrence in the Western United States Explained by Tropical and North Pacific Ocean Climate Variability. The American Journal of Tropical Medicine and Hygiene, 83(3), 624-632. Astrom, C., J. Rocklov, S. Hales, A. Beguin, V. Louis, and R. Sauerborn, 2012: Potential Distribution of Dengue Fever Under Scenarios of Climate Change and Economic Development. Ecohealth, 9(4), 448-454. Atchison, C., M. Iturriza-Gomara, C. Tam, and B. Lopman, 2010: Spatiotemporal dynamics of rotavirus disease in Europe: can climate or demographic variability explain the patterns observed. The Pediatric Infectious Disease Journal, 29(6), 566-568. Auffhammer, M., 2011: Agriculture: Weather dilemma for African maize. Nature Climate Change, 1, 27-28. Babey, S.H., T.A. Hastert, and E.R. Brown, 2007: Teens living in disadvantaged neighborhoods lack access to parks and get less physical activity. Policy Brief UCLA Cent Health Policy Res,(PB2007-4), 1-6. Babey, S.H., T.A. Hastert, H. Yu, and E.R. Brown, 2008: Physical activity among adolescents. When do parks matter? Am J Prev Med, 34(4), 345-8. Baccini, M., T. Kosatsky, A. Analitis, H.R. Anderson, M. D'Ovidio, B. Menne, P. Michelozzi, A. Biggeri, and P.C. Grp, 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. Bai, L., L.C. Morton, and Q. Liu, 2013: Climate change and mosquito-borne diseases in China: a review. Globalization and Health, 9(1), 10. Bandyopadhyay, S., S. Kanji, and L. Wang, 2012: The impact of rainfall and temperature variation on diarrheal prevalence in Sub-Saharan Africa. Applied Geography, 33(0), 63-72. Bangs, M.J., R.P. Larasati, A.L. Corwin, and S. Wuryadi, 2006: Climatic factors associated with epidemic dengue in Palembang, Indonesia: implications of short-term meteorological events on virus transmission. Southeast Asian J Trop Med Public Health, 37(6), 1103-16. Bassil, K. and D. Cole, 2010: Effectiveness of Public Health Interventions in Reducing Morbidity and Mortality during Heat Episodes: a Structured Review. International Journal of Environmental Research and Public Health, 7(3), 991-1001. Basu, R. and B.D. Ostro, 2008: A Multicounty Analysis Identifying the Populations Vulnerable to Mortality Associated with High Ambient Temperature in California. American Journal of Epidemiology, 168(6), 632-637. Battisti, D.S. and R.L. Naylor, 2009: Historical Warnings of Future Food Insecurity with Unprecedented Seasonal Heat. Science, 323(5911), 240-244. Baumgartner, J., J.J. Schauer, M. Ezzati, L. Lu, C. Cheng, J.A. Patz, and L.E. Bautista, 2011: Indoor Air Pollution and Blood Pressure in Adult Women Living in Rural China. Environmental Health Perspectives, 119(10), 1390-1395. Beaglehole, R. and R. Bonita, 2008: Global public health: a scorecard. Lancet, 372(9654), 1988-1996. 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 Negl Trop Dis, 3(5), e429. Beggs, P.J., 2010: Adaptation to impacts of climate change on aeroallergens and allergic respiratory diseases. International Journal of Environmental Research & Public Health, 7(8), 3006-3021. Béguin, A., S. Hales, J. Rocklöv, C. Åström, 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. Belanger, D., P. Gosselin, P. Valois, S. Germain, and B. Abdous, 2009: Use of a remote car starter in relation to smog and climate change perceptions: a population survey in Quebec (Canada). International Journal of Environmental Research & Public Health [Electronic Resource], 6(2), 694-709. Bell, M.L., R.D. Peng, and F. Dominici, 2006: The exposure-response curve for ozone and risk of mortality and the adequacy of current ozone regulations. Environ Health Perspect, 114(4), 532-6. Bell, M.L., D.L. Davis, L.A. Cifuentes, A.J. Krupnick, R.D. Morgenstern, and G.D. Thurston, 2008: Ancillary human health benefits of improved air quality resulting from climate change mitigation. Environmental Health, 7(41), 1-18. Bell, M., R. Goldberg, C. Hogrefe, P. Kinney, K. Knowlton, B. Lynn, J. Rosenthal, C. Rosenzweig, and J. Patz, 2007: Climate change, ambient ozone, and health in 50 US cities. Climatic Change, 82(1), 61-76. 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. Subject to Final Copyedit 39 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Bennet, L., A. Halling, and J. Berglund, 2006: Increased incidence of Lyme borreliosis in southern Sweden following mild winters and during warm, humid summers. Eur J Clin Microbiol Infect Dis, 25(7), 426-32. Bennett, C.M. and A.J. McMichael, 2010: Non-heat related impacts of climate change on working populations. Glob Health Action, 3(5460), 1-10. Bennett, C., K.G. Dear, and A. McMichael, 2013: Shifts in the seasonal distribution of deaths in Australia, 1968- 2007. International Journal of Biometeorology,(April), 1-8. Berry, H.L., 2007: "Crowded suburbs" and "killer cities": A brief review of the relationship between the urban environment and mental health. NSW Public Health Bulletin, 18(12), 222-227. 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), 123-132. Bhatt, S., P.W. Gething, O.J. Brady, J.P. Messina, A.W. Farlow, C.L. Moyes, J.M. Drake, J.S. Brownstein, A.G. Hoen, O. Sankoh, M.F. Myers, D.B. George, T. Jaenisch, G.R.W. Wint, C.P. Simmons, T.W. Scott, J.J. Farrar, and S.I. Hay, 2013: The global distribution and burden of dengue. Nature, 496(7446), 504-507. Bi, P., Y. Zhang, and K.A. Parton, 2007: Weather variables and Japanese encephalitis in the metropolitan area of Jinan city, China. J Infect, 55(6), 551-6. Bi, P. and 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? Aust J Rural Health, 16(1), 2-4. Black, R.E., L.H. Allen, Z.A. Bhutta, L.E. Caulfield, M.d. Onis, M. Ezzati, C. Mathers, and J. Rivera, 2008: Maternal and child undernutrition: global and regional exposures and health consequences. The Lancet, 371(9608), 243-260. Boldo, E., C. Linares, J. Lumbreras, R. Borge, A. Narros, J. Garcia-Perez, P. Fernandez-Navarro, B. Perez-Gomez, N. Aragones, R. Ramis, M. Pollan, T. Moreno, A. Karanasiou, and G. Lopez-Abente, 2010: Health impact assessment of a reduction in ambient PM(2.5) levels in Spain. Environ Int, 37(2), 342-8. Bompangue, N.D., P. Giraudoux, P.D. Plisnier, T.A. Mutombo, M. Piarroux, and B. Sudre, 2011: Dynamics of cholera outbreaks in Great Lakes region of Africa, 1978–2008. Emerg Infect Dis, 17(11), 2026-2034. Bond, T.C., S.J. Doherty, D.W. Fahey, P.M. Forster, T. Berntsen, B.J. DeAngelo, M.G. Flanner, S. Ghan, B. Kärcher, D. Koch, S. Kinne, Y. Kondo, P.K. Quinn, M.C. Sarofim, M.G. Schultz, M. Schulz, C. Venkataraman, H. Zhang, S. Zhang, N. Bellouin, S.K. Guttikunda, P.K. Hopke, M.Z. Jacobson, J.W. Kaiser, Z. Klimont, U. Lohmann, J.P. Schwarz, D. Shindell, T. Storelvmo, S.G. Warren, and C.S. Zender, 2013: Bounding the role of black carbon in the climate system: A scientific assessment. Journal of Geophysical Research: Atmospheres, 118(11), 5380-5552. Bonjour, S., H. Adair-Rohani, J. Wolf, S. Mehta, A. Pruss-Ustun, M. Lahiff, E.A. Rehfuess, V. Mishra, and K.R. Smith, 2013: Solid fuel use for household cooking: country and regional estimates for 1980-2010. Environmental Health Perspectives, 121(7), 784-790. Bonovas, S. and G. Nikolopoulos, 2012: High-burden epidemics in Greece in the era of economic crisis. Early signs of a public health tragedy. Journal of Preventive Medicine and Hygiene, 53(3), 169-171. Bouchama, A., M. Dehbi, G. Mohamed, F. Matthies, M. Shoukri, and B. Menne, 2007: Prognostic factors in heat wave related deaths: a meta-analysis. Archives of Internal Medicine, 167(20), 2170-2176. Bowen, K.J., S. Friel, K. Ebi, C.D. Butler, F. Miller, and A.J. McMichael, 2012: Governing for a healthy population: towards an understanding of how decision-making will determine our global health in a changing climate. International Journal of Environmental Research & Public Health [Electronic Resource], 9(1), 55-72. Brázdil, R., M. Trnka, P. Dobrovolný, K. Chromá, P. Hlavinka, and Z. Žalud, 2009: Variability of droughts in the Czech Republic, 1881–2006. Theoretical and Applied Climatology, 97(3-4), 297-315. Breton, M., M. Garneau, I. Fortier, F. Guay, and J. Louis, 2006: Relationship between climate, pollen concentrations of Ambrosia and medical consultations for allergic rhinitis in Montreal, 1994-2002. Science of the Total Environment, 370(1), 39-50. 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. Browning, C.R., D. Wallace, S.L. Feinberg, and K.A. Cagney, 2006: Neighborhood Social Processes, Physical Conditions, and Disaster-Related Mortality: The Case of the 1995 Chicago Heat Wave. American Sociological Review, 71(4), 661-678. Bruce, N., M. Dherani, J. Das, K. Balakrishnan, H. Adair-Rohani, Z. Bhutta, and D. Pope, 2013: Control of household air pollution for child survival: estimates for intervention impacts. BMC Public Health, 13, S8. Subject to Final Copyedit 40 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Brunkard, J., G. Namulanda, and R. Ratard, 2008: Hurricane Katrina deaths, Louisiana, 2005. Disaster Med Public Health Prep, 2(4), 215-223. Bujold, E., S.H. Mehta, C. Bujold, and R.J. Gauthier, 2002: Interdelivery interval and uterine rupture. Am J Obstet Gynecol, 187(5), 1199-202. Bulto, P., A.P. Rodriguez, A.R. Valencia, N.L. Vega, M.D. Gonzalez, and A.P. Carrera, 2006: Assessment of human health vulnerability to climate variability and change in Cuba. Environ Health Perspect, 114(12), 1942-1949. Byass, P., 2010: The Imperfect World of Global Health Estimates. PLoS Med, 7(11), e1001006. Caminade, C., J.M. Medlock, E. Ducheyne, K.M. McIntyre, S. Leach, M. Baylis, and A.P. Morse, 2012: Suitability of European climate for the Asian tiger mosquito Aedes albopictus: recent trends and future scenarios. Journal of the Royal Society Interface, 9(75), 2708-2717. Caminade, C., J.A. Ndione, C.M.F. Kebe, A.E. Jones, S. Danuor, S. Tay, Y.M. Tourre, J. 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. 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. Casagrande, S.S., M.C. Whitt-Glover, K.J. Lancaster, A.M. Odoms-Young, and T.L. Gary, 2009: Built environment and health behaviors among African Americans: a systematic review. Am J Prev Med, 36(2), 174-81. Casimiro, E., J. Calheiros, F.D. Santos, and R.S. Kovats, 2006: National assessment of health effects of climate change in Portugal: approach and key findings. Environ Health Perspect, 114(12), 1950-6. Chadee, D.D., B. Shivnauth, S.C. Rawlins, and A.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. 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(7), 2866-80. Chang, Y.K., C.C. Wu, L.T. Lee, R.S. Lin, Y.H. Yu, and Y.C. Chen, 2012: The short-term effects of air pollution on adolescent lung function in Taiwan. Chemosphere, 87(1), 26-30. Chapman, R.S., X. He, A.E. Blair, and Q. Lan, 2005: Improvement in household stoves and risk of chronic obstructive pulmonary disease in Xuanwei, China: retrospective cohort study. Bmj, 331(7524), 1050. Chau, P.H., K.C. Chan, and J. Woo, 2009: Hot weather warning might help to reduce elderly mortality in Hong Kong. International Journal of Biometeorology, 53(5), 461-468. Chaves, L.F. and C.J. Koenraadt, 2010: Climate change and highland malaria: fresh air for a hot debate. Q Rev Biol, 85(1), 27-55. Chaves, L.F., A. Satake, M. Hashizume, and N. Minakawa, 2012: Indian Ocean Dipole and Rainfall Drive a Moran Effect in East Africa Malaria Transmission. Journal of Infectious Diseases, 205(12), 1885-1891. Cheng, C.H., S.F. Huang, and H.J. Teoh, 2011: Predicting daily ozone concentration maxima using fuzzy time series based on a two-stage linguistic partition method. Computers & Mathematics with Applications, 62(4), 2016- 2028. Chowell, G., C.A. Torre, C. Munayco-Escate, L. Suarez-Ognio, R. Lopez-Cruz, J.M. Hyman, and C. Castillo- Chavez, 2008: Spatial and temporal dynamics of dengue fever in Peru: 1994-2006. Epidemiol Infect, 136(12), 1667-77. Chowell, G., B. Cazelles, H. Broutin, and C.V. Munayco, 2011: The influence of geographic and climate factors on the timing of dengue epidemics in Peru, 1994-2008. BMC Infect Dis, 11(164), 1-14. Christidis, N., P.A. Stott, G.S. Jones, H. Shiogama, T. Nozawa, and J. Luterbacher, 2012: Human activity and anomalously warm seasons in Europe. International Journal of Climatology, 32(2), 225-239. Cleary-Goldman, J., F.D. Malone, J. Vidaver, R.H. Ball, D.A. Nyberg, C.H. Comstock, G.R. Saade, K.A. Eddleman, S. Klugman, L. Dugoff, I.E. Timor-Tritsch, S.D. Craigo, S.R. Carr, H.M. Wolfe, D.W. Bianchi, and M. D'Alton, 2005: Impact of maternal age on obstetric outcome. Obstet Gynecol, 105(5 Pt 1), 983-90. Cohen, J.E., 2010: Population and climate change. Proceedings of the American Philosophical Society, 154(2), 158- 82. Conde-Agudelo, A., A. Rosas-Bermudez, and A.C. Kafury-Goeta, 2007: Effects of birth spacing on maternal health: a systematic review. Am J Obstet Gynecol, 196(4), 297-308. Constantin de Magny, G., R. Murtugudde, M.R.P. Sapiano, A. Nizam, C.W. Brown, A.J. Busalacchi, M. Yunus, G.B. Nair, A.I. Gil, C.F. Lanata, J. Calkins, B. Manna, K. Rajendran, M.K. Bhattacharya, A. Huq, R.B. Sack, Subject to Final Copyedit 41 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 and R.R. Colwell, 2008: Environmental signatures associated with cholera epidemics. Proceedings of the National Academy of Sciences, 105(46), 17676-17681. Cook, J.T. and D.A. Frank, 2008: Food Security, Poverty, and Human Development in the United States. Annals of the New York Academy of Sciences, 1136(1), 193-209. Cook, S.M., R.I. Glass, C.W. LeBaron, and M.S. Ho, 1990: Global seasonality of rotavirus infections. Bulletin of the World Health Organization, 68(2), 171-177. Correa, M.d.P., S. Godin-Beekmann, M. Haeffelin, S. Bekki, P. Saiag, J. Badosa, F. Jegou, A. Pazmino, and E. Mahe, 2013: Projected changes in clear-sky erythemal and vitamin D effective UV doses for Europe over the period 2006 to 2100. Photochemical & Photobiological Sciences, 12(6), 1053-1064. Danis, K., A. Baka, A. Lenglet, W. Van Bortel, I. Terzaki, M. Tseroni, M. Detsis, E. Papanikolaou, A. Balaska, S. Gewehr, G. Dougas, T. Sideroglou, A. Economopoulou, N. Vakalis, S. Tsiodras, S. Bonovas, and J. Kremastinou, 2011: Autochthonous Plasmodium vivax malaria in Greece, 2011. Euro Surveill, 16(42), 1-5. Dasgupta, S., B. Laplante, S. Murray, and D. Wheeler, 2009: Climate Change and the Future Impacts of Storm- Surge Disasters in Developing Countries. In: Working Paper 182. Center for Global Development, Washington, DC, pp. 1-24. Davis, T.W., D.L. Berry, G.L. Boyer, and C.J. Gobler, 2009: The effects of temperature and nutrients on the growth and dynamics of toxic and non-toxic strains of Microcystis during cyanobacteria blooms. Harmful Algae, 8(5), 715-725. De Haen, H. and G. Hemrich, 2007: The economics of natural disasters: implications and challenges for food security. Agricultural Economics, 37, 31-45. Dear, K., G. Ranmuthugala, T. Kjellstrom, C. Skinner, and I. Hanigan, 2005: Effects of temperature and ozone on daily mortality during the August 2003 heat wave in France. Archives of Environmental & Occupational Health, 60(4), 205-212. Delatte, H., G. Gimonneau, A. Triboire, and D. Fontenille, 2009: Influence of temperature on immature development, survival, longevity, fecundity, and gonotrophic cycles of Aedes albopictus, vector of chikungunya and dengue in the Indian Ocean. Journal of Medical Entomology, 46(1), 33-41. 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), 1225-33. Dennekamp, M. and M. Carey, 2010: Air quality and chronic disease: why action on climate change is also good for health. New South Wales Public Health Bulletin, 21(5-6), 115-21. Descloux, E., M. Mangeas, C.E. Menkes, M. Lengaigne, A. Leroy, T. Tehei, L. Guillaumot, M. Teurlai, A.C. Gourinat, J. Benzler, A. Pfannstiel, J.P. Grangeon, N. Degallier, and X. De Lamballerie, 2012: Climate-based models for understanding and forecasting dengue epidemics. PLoS Negl Trop Dis, 6(2), e1470. Dey, S.K., H. Ushijima, O. Phathammavong, W. Chanit, S. Okitsu, M. Mizuguchi, and Y. Ota, 2010: Seasonal trend and serotype distribution of rotavirus infection in Japan, 1981-2008. The Pediatric Infectious Disease Journal, 29(2), 166-167. Diamond-Smith, N. and M. Potts, 2011: A woman cannot die from a pregnancy she does not have. International Perspectives on Sexual and Reproductive Health, 37(3), 155-157. Diboulo, E., A. Sie, J. Rocklov, 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), 6-13. Diffenbaugh, N. and F. Giorgi, 2012: Climate change hotspots in the CMIP5 global climate model ensemble. Climatic Change, 114(3-4), 813-822. D'Ippoliti, D., P. Michelozzi, C. Marino, F. de'Donato, B. Menne, K. Katsouyanni, U. Kirchmayer, A. Analitis, M. Medina-Ramón, A. Paldy, R. Atkinson, S. Kovats, L. Bisanti, A. Schneider, A. Lefranc, C. Iñiguez, and C.A. Perucci, 2010: The impact of heat waves on mortality in 9 European cities : results from the EuroHEAT project. Environmental Health, 9(37), 1-9. Dix-Cooper, L., B. Eskenazi, C. Romero, J. Balmes, and K.R. Smith, 2012: Neurodevelopmental performance among school age children in rural Guatemala is associated with prenatal and postnatal exposure to carbon monoxide, a marker for exposure to woodsmoke. Neurotoxicology, 33(2), 246-254. Dodman, D., D. Mitlin, and J.C. Rayos Co, 2010: Victims to victors, disasters to opportunities: community-driven responses to climate change in the Philippines. International Development Planning Review, 32(1), 1-26. Subject to Final Copyedit 42 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Dong, G.H., T. Chen, M.M. Liu, D. Wang, Y.N. Ma, W.H. Ren, Y.L. Lee, Y.D. Zhao, and Q.C. He, 2011: Gender differences and effect of air pollution on asthma in children with and without allergic predisposition: northeast Chinese children health study. PLoS One, 6(7), e22470. Doocy, S., A. Daniels, S. Murray, and T.D. Kirsch, 2013: The Human Impact of Floods: a Historical Review of Events 1980-2009 and Systematic Literature Review. PLOS Currents Disasters, 1, 1-21. Dossou, K. and B. Glehouenou-Dossou, 2007: The vulnerability to climate change of Cotonou (Benin): the rise in sea level. Environment and Urbanization, 19, 65-79. Doyon, B., D. Belanger, and P. Gosselin, 2008: The potential impact of climate change on annual and seasonal mortality for three cities in Quebec, Canada. Int J Health Geogr, 7(23), 1-12. Dunne, J., R. Stouffer, and J. John, 2013: Reductions in labour capacity from heat stress under climate warming. Nature Clim. Change, 3(3), 1-4. Durand, C.P., M. Andalib, G.F. Dunton, J. Wolch, and M.A. Pentz, 2011: A systematic review of built environment factors related to physical activity and obesity risk: implications for smart growth urban planning. Obes Rev, 12(5), e173-82. Earnest, A., S.B. Tan, and A. Wilder-Smith, 2012: Meteorological factors and El Nino Southern Oscillation are independently associated with dengue infections. Epidemiol Infect, 140(7), 1244-1251. Ebi, K., J. Smith, I. Burton, and J. Scheraga, 2006: Some lessons learned from public health on the process of adaptation. Mitigation Adaptation Strategies Global Change, 11, 607-620. Ebi, K.L., K.A. Exuzides, E. Lau, M. Kelsh, and A. Barnston, 2004: Weather changes associated with hospitalizations for cardiovascular diseases and stroke in California, 1983-1998. International Journal of Biometeorology, 49(1), 48-58. Ebi, K.L. and G. McGregor, 2008: Climate change, tropospheric ozone and particulate matter, and health impacts. Environ Health Perspect, 116(11), 1449-55. Ebi, K.L. and D. Mills, 2013: Winter mortality in a warming world: a re-assessment. WIREs Climate Change, 4, 203-212. Edwards, R.D., K.R. Smith, J. Zhang, and Y. Ma, 2004: Implications of changes in household stoves and fuel use in China. Energy Policy, 32(3), 395-411. Engelman, R., 2010: Population, Climate Change, and Women's Lives. In: Worldwatch Reports [Mastny, L. (ed.)]. Worldwatch, Washington, D.C., pp. 183. Erdner, D.L., J. Dyble, M.L. Parsons, R.C. Stevens, K.A. Hubbard, M.L. Wrabel, S.K. Moore, K.A. Lefebvre, D.M. Anderson, P. Bienfang, R.R. Bidigare, M.S. Parker, P. Moeller, L.E. Brand, and V.L. Trainer, 2008: Centers for Oceans and Human Health: a unified approach to the challenge of harmful algal blooms. Environmental Health: A Global Access Science Source, 7 Suppl 2, S2. Estrada-Peña, A., N. Ayllón, and J. de la Fuente, 2012: Impact of climate trends on tick-borne pathogen transmission. Front Physiol, 3, 64. Ezzine-de-Blas, D., J. Börner, A. Violato-Espada, N. Nascimento, and M. Piketty, 2011: Forest loss and management in land reform settlements: Implications for REDD governance in the Brazilian Amazon. Environmental Science & Policy, 14(2), 188-200. Fang, L.Q., X.J. Wang, S. Liang, Y.L. Li, S.X. Song, W.Y. Zhang, Q. Qian, Y.P. Li, L. Wei, Z.Q. Wang, H. Yang, and W.C. Cao, 2010: Spatiotemporal trends and climatic factors of hemorrhagic fever with renal syndrome epidemic in Shandong Province, China. PLoS Negl Trop Dis, 4(8), e789. Feachem, R.G., 2011: The Global Fund: getting the reforms right. The Lancet, 378(9805), 1764-1765. Finlay, S., A. Moffat, R. Gazzard, D. Baker, and V. Murray, 2012: Health impacts of wildfires. PLoS Currents,. Finucane, M.M., G.A. Stevens, M.J. Cowan, G. Danaei, J.K. Lin, C.J. Paciorek, G.M. Singh, H.R. Gutierrez, Y.L. Msc, A.N.B. Meng, F.F. Md, L.M.R. Msc, P.M.E. Phd, and Index,on behalf of the Global Burden of Metabolic Risk Factors of Chronic Diseases Collaborating Group Body Mass, 2011: National, regional, and global trends in body-mass index since 1980: systematic analysis of health examination surveys and epidemiological studies with 960 country-years and 9.1 million participants. The Lancet, 377(9765), 557-567. Foley, J.A., N. Ramankutty, K.A. Brauman, E.S. Cassidy, J.S. Gerber, M. Johnston, N.D. Mueller, C. O’Connell, D.K. Ray, P.C. West, C. Balzer, E.M. Bennett, S.R. Carpenter, J. Hill, C. Monfreda, S. Polasky, J. Rockström, J. Sheehan, S. Siebert, D. Tilman, and D.P.M. Zaks, 2011: Solutions for a cultivated planet. Nature, 337(7369), 337-342. Ford, J.D., T. Pearce, J. Gilligan, B. Smit, and J. Oakes, 2008: Climate change and hazards associated with ice use in northern Canada. Arctic, Antarctic, and Alpine Research, 40(4), 647-659. Subject to Final Copyedit 43 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Ford, J.D., 2009: Vulnerability of Inuit food systems to food insecurity as a consequence of climate change: a case study from Igloolik, Nunavut. Reg Environ Change, 9(2), 83-100. Fouillet, A., G. Rey, F. Laurent, G. Pavillon, S. Bellec, C. Guihenneuc-Jouyaux, J. Clavel, E. Jougla, and D. Hémon, 2006: Excess mortality related to the August 2003 heat wave in France. International Archives of Occupational & Environmental Health, 80(1), 16-24. Fouillet, A., G. Rey, V. Wagner, K. Laaidi, P. Empereur-Bissonnet, A. Le Tertre, P. Frayssinet, P. Bessemoulin, F. Laurent, P. De Crouy-Chanel, E. Jougla, and D. Hémon, 2008: Has the impact of heat waves on mortality changed in France since the European heat wave of summer 2003? A study of the 2006 heat wave. International Journal of Epidemiology, 37(2), 309-317. Friel, S., M. Marmot, A.J. McMichael, T. Kjellstrom, and D. Vagero, 2008: Global health equity and climate stabilisation: a common agenda. Lancet, 372(9650), 1677-83. Friel, S., A.D. Dangour, T. Garnett, K. Lock, Z. Chalabi, I. Roberts, A. Butler, C.D. Butler, J. Waage, A.J. McMichael, and A. Haines, 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: food and agriculture. Lancet, 374(9706), 2016-25. Frumkin, H., J. Hess, G. Luber, J. Malilay, and M. McGeehin, 2008: Climate Change: The Public Health Response. American Journal of Public Health, 98(3), 435-445. Frumkin, H. and A.J. McMichael, 2008: Climate change and public health: thinking, communicating, acting. American Journal of Preventive Medicine, 35(5), 403-10. Gamble, J.L., B.J. Hurley, P.A. Schultz, W.S. Jaglom, N. Krishnan, and M. Harris, 2013: Climate change and older Americans: state of the science. Environmental Health Perspectives, 121(1), 15-22. Gasparrini, A. and B. Armstrong, 2011: The impact of heat waves on mortality. Epidemiology (Cambridge, Mass.), 22(1), 68-73. Ge, C., C. Lee, and J. Lee, 2012: The impact of extreme weather events on Salmonella internalization in lettuce and green onion. Food Research International, 45(2), 1118-1122. GEA, 2012: Global Energy Assessment - Toward a Sustainable Future. Cambridge University Press and the International Institute for Applied Systems Analysis, Laxenburg, Austria., Cambridge, UK and New York, NY, USA, pp. 1-1884. Gemenne, F., 2011: Climate-induced population displacements in a 4°C+ world. Phil.Trans.R.Soc.A, 369(1934), 182-195. 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. 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 Infect Dis, 11(166), 1-13. Gosling, S.N., G.R. Mcgregor, and 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. Grabow, M.L., S.N. Spak, T. Holloway, B. Stone Jr., A.C. Mednick, and J.A. Patz, 2011: Air Quality and Exercise- Related Health Benefits from Reduced Car Travel in the Midwestern United States. Environ Health Perspect, 120(1), 68-76. 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(1–2), 405-413. Gray, J.S., H. Dautel, A. Estrada-Pena, O. Kahl, and E. Lindgren, 2009: Effects of climate change on ticks and tick- borne diseases in Europe. Interdiscip Perspect Infect Dis, 2009(593232), 1-12. Green, R., L. Cornelsen, A.D. Dangour, R. Turner, B. Shankar, M. Mazzocchi, and R.D. Smith, 2013: The effect of rising food prices on food consumption: systematic review with meta-regression. Bmj, 346. Gribble, J.N., N.J. Murray, and E.P. Menotti, 2009: Reconsidering childhood undernutrition: can birth spacing make a difference? An analysis of the 2002-2003 El Salvador National Family Health Survey. Matern Child Nutr, 5(1), 49-63. Gubler, D.J. and G. Kuno, 1997: Dengue and Dengue Haemorrhagic Fever. CAB International, Cambridge. Guha-Sapir, D., F. Vos, and R. Below, 2011: Annual Disaster Statistical Review 2010 The numbers and trends. Center for Research on the Epidemiology of Disasters, Brussels, pp. 1-50. Ha, J., H. Kim, and S. Hajat, 2011: Effect of Previous-Winter Mortality on the Association between Summer Temperature and Mortality in South Korea. Environmental Health Perspectives, 119(4), 542-546. Subject to Final Copyedit 44 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Haines, A., K.R. Smith, D. Anderson, P.R. Epstein, A.J. McMichael, I. Roberts, P. Wilkinson, J. Woodcock, and J. Woods, 2007: Policies for accelerating access to clean energy, improving health, advancing development, and mitigating climate change. Lancet, 370(9594), 1264-81. Haines, A., A. McMichael, K. Smith, I. Roberts, J. Woodcock, A. Markandya, B. Armstrong, D. Campbell- Lendrum, A. Dangour, M. Davies, N. Bruce, C. Tonne, M. Barrett, and P. Wilkinson, 2009: Health and Climate Change 6: Public health benefits of strategies to reduce greenhouse-gas emissions: overview and implications for policy makers. Lancet, 374(9707), 2104-14. 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. Halide, H. and P. Ridd, 2008: A predictive model for Dengue Hemorrhagic Fever epidemics. Int J Environ Health Res, 18(4), 253-65. Hancock, P., J. Ross, and J. Szalma, 2007: A meta-analysis of performance response under thermal stressors. Human Factors, 49, 851-877. Handmer, J., Y. Honda, Z.W. Kundzewicz, N. Arnell, G. Benito, J. Hatfield, I.F. Mohamed, P. Peduzzi, S. Wu, B. Sherstyukov, and K. Takahashi, 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. [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi et al. (eds.)]. Cambridge University Press., Cambridge, pp. 231-290. Hanigan, I.C., C.D. Butler, P.N. Kokic, and 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. 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 Pac J Public Health, 23(2 Suppl), 14S-26. Hansen, A.L., P. Bi, P. Ryan, M. Nitschke, D. Pisaniello, and G. Tucker, 2008: The effect of heat waves on hospital admissions for renal disease in a temperate city of Australia. Int J Epidemiol, 37(6), 1359-65. Hashizume, M. and A.M. Dewan, 2012: Hydroclimatological variability and dengue transmission in Dhaka, Bangladesh: a time-series study. BMC Infect Dis, 12(1), 98. Hashizume, M., 2008: Factors determining vulnerability to diarrhoea during and after severe floods in Bangladesh. Journal of Water and Health, 6(3), 323. Health Protection Agency, 2012: Health Effects of Climate Change in the UK 2012: Current evidence, recommendations and research gaps. [Vardoulakis, S. and Heaviside, C. (eds.)]. Centre for Radiation, Chemical and Environmental Hazards, Chilton, UK, pp. 1-242. Heguy, L., M. Garneau, M.S. Goldberg, M. Raphoz, F. Guay, and M. Valois, 2008: Associations between grass and weed pollen and emergency department visits for asthma among children in Montreal. Environmental Research, 106(2), 203-211. Herrera-Martinez, A.D. and A.J. Rodriguez-Morales, 2010: Potential influence of climate variability on dengue incidence registered in a western pediatric Hospital of Venezuela. Trop Biomed, 27(2), 280-6. Hess, J.J., 2008: Climate Change: The Importance of Place. American Journal of Preventive Medicine, 35(5), 468. Hesterberg, T.W., W.B. Bunn, R.O. McClellan, A.K. Hamade, C.M. Long, and P.A. Valberg, 2009: Critical review of the human data on short-term nitrogen dioxide (NO2) exposures: evidence for NO2 no-effect levels. Crit Rev Toxicol, 39(9), 743-81. Hii, Y.L., J. Rocklov, 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. Glob Health Action, 2, 1-9. Hii, Y.L., J. Rocklov, and N. Ng, 2011: Short term effects of weather on hand, foot and mouth disease. PLoS One, 6(2), e16796. Hii, Y.L., H. Zhu, N. Ng, L.C. Ng, and J. Rocklov, 2012: Forecast of Dengue Incidence Using Temperature and Rainfall. PLoS Negl Trop Dis, 6(11), e1908. 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. Holt, A.C., D.J. Salkeld, C.L. Fritz, J.R. Tucker, and P. Gong, 2009: Spatial analysis of plague in California: niche modeling predictions of the current distribution and potential response to climate change. Int J Health Geogr, 8(38), 1-14. Subject to Final Copyedit 45 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Honda, Y., M. Kondo, G. McGregor, H. Kim, Y. Guo, Y. Hijioka, M. Yoshikawa, K. Oka, S. Takano, S. Hales, and R.S. Kovats, 2013: Heat-related mortality risk model for climate change impact projection. Environmental Health and Preventive Medicine, 1-8. Hondula, D.M., J. Rocklov, and O.A. Sankoh, 2012: Past, present, and future climate at select INDEPTH member Health and Demographic Surveillance Systems in Africa and Asia. Glob Health Action, 5, 74-86. Hooper, L., C.D. Summerbell , R. Thompson, D. Sills, F.G. Roberts, H.J. Moore, and G. Davey Smith, 2012: Reduced or modified dietary fat for preventing cardiovascular disease. Cochrane Database of Systematic Reviews, 5, 1-223. Hosking, J. and D. Campbell-Lendrum, 2012: How well does climate change and human health research match the demands of policymakers? A scoping review. Environmental Health Perspectives, 120(8), 1076-82. Hsieh, Y.H. and C.W. Chen, 2009: Turning points, reproduction number, and impact of climatological events for multi-wave dengue outbreaks. Trop Med Int Health, 14(6), 628-38. 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), 1681-90. Huang, C., A.G. Barnett, X. Wang, and S. Tong, 2012: The impact of temperature on years of life lost in Brisbane, Australia. Nature Climate Change, 2(4), 265-270. Hughes, B.B., R. Kuhn, C.M. Peterson, D.S. Rothman, J.R. Solorzano, C.D. Mathers, and J.R. Dickson, 2011: Projections of global health outcomes from 2005 to 2060 using the International Futures integrated forecasting model. Bull World Health Organ, 89(7), 478-86. International Fund for Agricultural Development, 2010: Rural Poverty Report 2011: New realities, new challenges: new opportunities for tomorrow’s generation. International Fund for Agricultural Development, Rome, pp. 1- 322. International Labor Organization, 2013: Global Employment Trends 2013: Recovering from a second jobs dip. International Labor Organization, Geneva, pp. 1-239. IPCC, 2007: Working Group I Report: "The Physical Science Basis". Intergovernmental Panel on Climate Change. 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, UK, and New York, NY, USA, pp. 582. Islam, M.S., 2009: Effects of local climate variability on transmission dynamics of cholera in Matlab, Bangladesh. Transactions of the Royal Society of Tropical Medicine and Hygiene, 103(11), 1165. ISO, 1989: Hot environments - Estimation of the heat stress on working man, based on the WBGT-index (wet bulb globe temperature). ISO Standard 7243. International Organization for Standardization, Geneva. 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. Climate Change, 102(1-2), 159-186. Jacobson, M.Z., 2009: Review of solutions to global warming, air pollution, and energy security. Energy & Environmental Science,(2), 148-173. Jacobson, M.Z., 2010: Short-term effects of controlling fossil-fuel soot, biofuel soot and gases, and methane on climate, Arctic ice, and air pollution health. Journal of Geophysical Research: Atmospheres, 115(D14), - D14209. Jacobson, M.Z., R.W. Howarth, M.A. Delucchi, S.R. Scobie, J.M. Barth, M.J. Dvorak, M. Klevze, H. Katkhuda, B. Miranda, N.A. Chowdhury, R. Jones, L. Plano, and A.R. Ingraffea, 2013: Examining the feasibility of converting New York State’s all-purpose energy infrastructure to one using wind, water, and sunlight. Energy Policy, 57(0), 585-601. Jaenson, T., D. Jaenson, L. Eisen, E. Petersson, and E. Lindgren, 2012: Changes in the geographical distribution and abundance of the tick Ixodes ricinus during the past 30 years in Sweden. Parasites & Vectors, 5(1), 8. Jakszyn, P., C.A. González, L. Luján-Barroso, M.M. Ros, H.B. Bueno-de-Mesquita, N. Roswall, A.M. Tjønneland, F.L. Büchner, L. Egevad, K. Overvad, O. Raaschou-Nielsen, F. Clavel-Chapelon, M. Boutron-Ruault, M.S. Touillaud, J. Chang-Claude, N.E. Allen, L.A. Kiemeney, T.J. Key, R. Kaaks, H. Boeing, S. Weikert, A. Trichopoulou, E. Oikonomou, D. Zylis, D. Palli, F. Berrino, P. Vineis, R. Tumino, A. Mattiello, P.H.M. Peeters, C.L. Parr, I.T. Gram, G. Skeie, M. Sánchez, N. Larrañaga, E. Ardanaz, C. Navarro, L. Rodríguez, D. Ulmert, R. Ehrnström, G. Hallmans, B. Ljungberg, A.W. Roddam, S.A. Bingham, K. Khaw, N. Slimani, P.A. Boffetta, M. Jenab, T. Mouw, D.S. Michaud, and E. Riboli, 2011: Red Meat, Dietary Nitrosamines, and Heme Iron and Risk Subject to Final Copyedit 46 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 of Bladder Cancer in the European Prospective Investigation into Cancer and Nutrition (EPIC). Cancer Epidemiology Biomarkers & Prevention, 20(3), 555-559. Jakubicka, T., F. Vos, R. Phalkey, M. Marx, and D. Guha-Sapir, 2010: Health impacts of floods in Europe: data gaps and information needs from a spatial perspective. Université Catholique de Louvain. Institut de Recherche Santé et Société (UCL), Microdis, Louvain, pp. 1-44. Jamison, D.T., J.G. Breman, A.R. Measham, G. Alleyne, M. Claeson, D.B. Evans, J. Prabhat, A. Mills, and P. Musgrove (eds.), 2006: Disease Control Priorities in Developing Countries The World Bank and Oxford University Press., New York, pp. 1400. Jankowska, M.M., D. Lopez-Carr, C. Funk, and Z.A. Chafe, 2012: Climate change and human health: Spatial modeling availability, malnutrition, and livelihoods in Mali, Africa. Applied Geography, 33, 4-15. Jarrett, J., J. Woodcock, U.K. Griffiths, Z. Chalabi, P. Edwards, I. Roberts, and A. Haines, 2012: Effect of increasing active travel in urban England and Wales on costs to the National Health Service. The Lancet, 379(9832), 2198-2205. Jensen, H., M. Keogh-Brown, R. Smith, Z. Chalabi, A. Dangour, M. Davies, P. Edwards, T. Garnett, M. Givoni, U. Griffiths, I. Hamilton, J. Jarrett, I. Roberts, P. Wilkinson, J. Woodcock, and A. Haines, 2013: The importance of health co-benefits in macroeconomic assessments of UK Greenhouse Gas emission reduction strategies. Climatic Change, 1-15. Jerrett, M., R.T. Burnett, C.A. Pope 3rd, K. Ito, G. Thurston, D. Krewski, Y. Shi, E. Calle, and M. Thun, 2009: Long-term ozone exposure and mortality. N Engl J Med, 360(11), 1085-95. Johnston, F., S. Henderson, Y. Chen, J. Randerson, M. Marlier, R. DeFries, P. Kinney, D. Bowman, and B. M, 2012: Estimated global mortality attributable to smoke from landscape fires. Environmental Health Perspectives, 120(5), 695-701. Jonkman, S.N. and I. Kelman, 2005: An analysis of the causes and circumstances of flood disaster deaths. Disasters, 29(1), 75-97. Kaczynski, A.T. and K.A. Henderson, 2008: Parks and recreation settings and active living: a review of associations with physical activity function and intensity. J Phys Act Health, 5(4), 619-32. Karanikolos, M., P. Mladovsky, J. Cylus, S. Thomson, S. Basu, D. Stuckler, J.P. Mackenbach, and M. McKee, 2013: Financial crisis, austerity, and health in Europe. The Lancet, 381(9874), 1323-1331. Kearney, M., W.P. Porter, C. Williams, S. Ritchie, and 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. 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. Kellenberg, D.K. and A.M. Mobarak, 2008: Does rising income increase or decrease damage risk from natural disasters? Journal of Urban Economics, 63(3), 788-802. 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. Malar J, 8(268), 1-8. Kessler, R.C., S. Galea, M.J. Gruber, N.A. Sampson, R.J. Ursano, and S. Wessely, 2008: Trends in mental illness and suicidality after Hurricane Katrina. Molecular Psychiatry, 13(4), 374-384. Khan, M.S.A., 2008: Disaster preparedness for sustainable development in Bangladesh. Disaster Prevention and Management, 17(5), 662-671. Khetsuriani, N., A. LaMonte-Fowlkes, M.S. Oberste, and M.A. Pallansch, 2006: Enterovirus surveillance—United States, 1970–2005. MMWR. Surveillance Summaries, 55(8), 1. Kim, Y., H. Kim, and D.S. Kim, 2011: Association between daily environmental temperature and suicide mortality in Korea (2001-2005). Psychiatry Res, 186(2-3), 390-6. Kinney, P.L., M. Pascal, R. Vautard, and K. Laaidi, 2012: La mortalite hivernale va-t-elle diminuer avec le changement climatique? [Winter mortality in a changing climate: will it go down?]. Bull Epidemiol Hebd,(12- 13), 149-151. Kinney, P.L., M.S. O’Neill, M.L. 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. Kite-Powell, H.L., L.E. Fleming, L.C. Backer, E.M. Faustman, P. Hoagland, A. Tsuchiya, L.R. Younglove, B.A. Wilcox, and R.J. Gast, 2008: Linking the oceans to public health: current efforts and future directions. Environmental Health, 7(Suppl 2), S6. Subject to Final Copyedit 47 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Kjellstrom, T., B. Lemke, and O. Hyatt, 2011: Increased workplace heat exposure due to climate change. Asia- Pacific Newsletter on Occupational Health and Safety,(18), 6-20. Kjellstrom, T., B. Lemke, and M. Otto, 2013: Mapping occupational heat exposure and effects in South-East Asia: Ongoing time trends 1980-2009 and future estimates to 2050. Industrial Health, 51, 56-67. Kjellstrom, T., I. Holmer, and B. Lemke, 2009a: Workplace heat stress, health and productivity - an increasing challenge for low and middle-income countries during climate change. Glob Health Action, 2(1), 6. Kjellstrom, T., R.S. Kovats, S.J. Lloyd, T. Holt, and R.S. Tol, 2009b: The direct impact of climate change on regional labor productivity. Arch Environ Occup Health, 64(4), 217-27. Kjellstrom, T. and J. Crowe, 2011: Climate change, workplace heat exposure, and occupational health and productivity in Central America. International Journal of Occupational & Environmental Health, 17(3), 270- 81. Knowlton, K., B. Lynn, R.A. Goldberg, C. Rosenzweig, C. Hogrefe, J.K. Rosenthal, and P.L. Kinney, 2007: Projecting heat-related mortality impacts under a changing climate in the New York City region. American Journal of Public Health, 97(11), 2028-2034. 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. 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. Kolstad, E. and K.A. Johansson, 2011: Uncertainties associated with quantifying climate change impacts on human health: a case study for diarrhea. Environ Health Perspect, 119(3), 299-305. Kosatsky, T., 2005: The 2003 European heat waves. Euro Surveill, 10(7), 148-149. Kovats, R.S., T. Wolf, and B. Menne, 2004: Heatwave of August 2003 in Europe: provisional estimates of the impact on mortality. EuroSurveillance Weekly, 8(11). Kovats, R.S. and S. Hajat, 2008: Heat stress and public health: a critical review. Annu Rev Public Health, 29, 41-55. Kozuki, N., A. Lee, M. Silveira, C. Victora, L. Adair, J. Humphrey, R. Ntozini, R. Black, J. Katz, and Child Health Epidemiology Reference Group Small-for-Gestational-Age-Preterm Birth,Working Group, 2013: The associations of birth intervals with small-for-gestational-age, preterm, and neonatal and infant mortality: a meta- analysis. BMC Public Health, 13, S3. Kriz, B., M. Maly, C. Benes, and M. Daniel, 2012: Epidemiology of Tick-Borne Encephalitis in the Czech Republic 1970-2008. Vector Borne Zoonotic Dis, 12(11), 994-9. Lai, L.W., 2011: Influence of environmental conditions on asynchronous outbreaks of dengue disease and increasing vector population in Kaohsiung, Taiwan. International Journal of Environmental Health Research, 21, 133- 146. Lake, I.R., 2009: A re-evaluation of the impact of temperature and climate change on foodborne illness. Epidemiology and Infection, 137(11), 1538. Lake, I.R., L. Hooper, A. Abdelhamid, G. Bentham, A.B.A. Boxall, A. Draper, S. Fairweather-Tait, M. Hulme, P.R. Hunter, G. Nichols, and K.W. Waldron, 2012: Climate Change and Food Security: Health Impacts in Developed Countries. Environmental Health Perspectives, 120(11), 1520-1526. Lam, N.L., Y. Chen, C. Weyant, C. Venkataraman, P. Sadavarte, M.A. Johnson, K.R. Smith, B.T. Brem, J. Arineitwe, J.E. Ellis, and T.C. Bond, 2012: Household Light Makes Global Heat: High Black Carbon Emissions From Kerosene Wick Lamps. Environmental Science & Technology, 46(24), 13531-13538. Lan, Q., R.S. Chapman, D.M. Schreinemachers, L. Tian, and X. He, 2002: Household stove improvement and risk of lung cancer in Xuanwei, China. J Natl Cancer Inst, 94(11), 826-35. Lecocq, F. and Z. Shalizi, 2007: Balancing expenditures on mitigation of and adaptation to climate change: an exploration of Issues relevant to developing countries. In: Policy Research Working Paper. The World Bank, Washington, DC, pp. 1-48. Lefohn, A.S., D. Shadwick, and S.J. Oltmans, 2010: Characterizing changes in surface ozone levels in metropolitan and rural areas in the United States for 1980-2008 and 1994-2008. Atmospheric Environment, 44(39), 5199- 5210. Lemke, B. and T. Kjellstrom, 2012: Calculating workplace WBGT from meteorological data: a tool for climate change assessment. Industrial Health, 50(4), 267-278. Subject to Final Copyedit 48 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Lepeule, J., F. Laden, D. Dockery, and J. Schwartz, 2012: Chronic Exposure to Fine Particles and Mortality: An Extended Follow-Up of the Harvard Six Cities Study from 1974 to 2009. Environ Health Perspect, 120(7), 965- 970. Li, S., H. Tao, and Y. Xu, 2011: Abiotic Determinants to the Spatial Dynamics of Dengue Fever in Guangzhou. Asia Pac J Public Health, 25(3), 239-247. Likhvar, V., Y. Honda, and M. Ono, 2011: Relation between temperature and suicide mortality in Japan in the presence of other confounding factors using time-series analysis with a semiparametric approach. Environ Health Prev Med, 16(1), 36-43. Lim, S.S., T. Vos, A.D. Flaxman, G. Danaei, K. Shibuya, H. Adair-Rohani, M.A. AlMazroa, M. Amann, H.R. Anderson, K.G. Andrews, M. Aryee, C. Atkinson, L.J. Bacchus, A.N. Bahalim, K. Balakrishnan, J. Balmes, S. Barker-Collo, A. Baxter, M.L. Bell, J.D. Blore, F. Blyth, C. Bonner, G. Borges, R. Bourne, M. Boussinesq, M. Brauer, P. Brooks, N.G. Bruce, B. Brunekreef, C. Bryan-Hancock, C. Bucello, R. Buchbinder, F. Bull, R.T. Burnett, T.E. Byers, B. Calabria, J. Carapetis, E. Carnahan, Z. Chafe, F. Charlson, H. Chen, J.S. Chen, A.T. Cheng, J.C. Child, A. Cohen, K.E. Colson, B.C. Cowie, S. Darby, S. Darling, A. Davis, L. Degenhardt, F. Dentener, D. Des Jarlais C., K. Devries, M. Dherani, E.L. Ding, E.R. Dorsey, T. Driscoll, K. Edmond, S.E. Ali, R.E. Engell, P.J. Erwin, S. Fahimi, G. Falder, F. Farzadfar, A. Ferrari, M.M. Finucane, S. Flaxman, F.G.R. Fowkes, G. Freedman, M.K. Freeman, E. Gakidou, S. Ghosh, E. Giovannucci, G. Gmel, K. Graham, R. Grainger, B. Grant, D. Gunnell, H.R. Gutierrez, W. Hall, H.W. Hoek, A. Hogan, H.D. Hosgood, D. Hoy, H. Hu, B.J. Hubbell, S.J. Hutchings, S.E. Ibeanusi, G.L. Jacklyn, R. Jasrasaria, J.B. Jonas, H. Kan, J.A. Kanis, N. Kassebaum, N. Kawakami, Y. Khang, S. Khatibzadeh, J. Khoo, C. Kok, F. Laden, R. Lalloo, Q. Lan, T. Lathlean, J.L. Leasher, J. Leigh, Y. Li, J.K. Lin, S.E. Lipshultz, S. London, R. Lozano, Y. Lu, J. Mak, R. Malekzadeh, L. Mallinger, W. Marcenes, L. March, R. Marks, R. Martin, P. McGale, J. McGrath, S. Mehta, Z.A. Memish, G.A. Mensah, T.R. Merriman, R. Micha, C. Michaud, V. Mishra, K.M. Hanafiah, A.A. Mokdad, L. Morawska, D. Mozaffarian, T. Murphy, M. Naghavi, B. Neal, P.K. Nelson, J.M. Nolla, R. Norman, C. Olives, S.B. Omer, J. Orchard, R. Osborne, B. Ostro, A. Page, K.D. Pandey, C.D.H. Parry, E. Passmore, J. Patra, N. Pearce, P.M. Pelizzari, M. Petzold, M.R. Phillips, D. Pope, C.A. Pope, J. Powles, M. Rao, H. Razavi, E.A. Rehfuess, J. Rehm, B. Ritz, F.P. Rivara, T. Roberts, C. Robinson, J. Rodriguez-Portales, I. Romieu, R. Room, L.C. Rosenfeld, A. Roy, L. Rushton, J.A. Salomon, U. Sampson, L. Sanchez-Riera, E. Sanman, A. Sapkota, S. Seedat, P. Shi, K. Shield, R. Shivakoti, G.M. Singh, D.A. Sleet, E. Smith, K.R. Smith, N.J.C. Stapelberg, K. Steenland, H. Stöckl, L.J. Stovner, K. Straif, L. Straney, G.D. Thurston, J.H. Tran, R. Van Dingenen, A. van Donkelaar, J.L. Veerman, L. Vijayakumar, R. Weintraub, M.M. Weissman, R.A. White, H. Whiteford, S.T. Wiersma, J.D. Wilkinson, H.C. Williams, W. Williams, N. Wilson, A.D. Woolf, P. Yip, J.M. Zielinski, A.D. Lopez, C.J.L. Murray, and M. Ezzati, 2012: A comparative risk assessment of burden of disease and injury attributable to 67 risk factors and risk factor clusters in 21 regions, 1990-2010: a systematic analysis for the Global Burden of Disease Study 2010. Lancet, 380(9859), 2224-2260. Lin, R.T. and C.C. Chan, 2009: Effects of heat on workers' health and productivity in Taiwan. Glob Health Action, 2. Lindgren, E. and R. Gustafson, 2001: Tick-borne encephalitis in Sweden and climate change. Lancet, 358(9275), 16-18. Liu, X., B. Jiang, W. Gu, and Q. Liu, 2011: Temporal trend and climate factors of hemorrhagic fever with renal syndrome epidemic in Shenyang City, China. Bmc Infectious Diseases, 11, 331. Liu, C., N. Hofstra, and E. Franz, 2013a: Impacts of climate change on the microbial safety of pre-harvest leafy green vegetables as indicated by Escherichia coli O157 and Salmonella spp. International Journal of Food Microbiology, 163(2), 119-128. Liu, Q., X. Liu, Cirendunzhu, A. Woodward, Pengcuociren, L. Bai, Baimaciwang, S. Sang, Dazhen, F. Wan, L. Zhou, Y. Guo, H. Wu, G. Li, L. Lu, J. Wang, Dawa, C. Chu, and Xiraoruodeng, 2013b: Mosquitoes established in Lhasa city, Tibet, China. Parasites and Vectors, 6(224), 1-10. Lloyd, S.J., R.S. Kovats, and Z. Chalabi, 2011: Climate change, crop yields, and malnutrition: development of a model to quantify the impact of climate scenarios on child malnutrition. Environmental Health Perspectives, 119(12), 1817-1823. Lobell, D.B., W. Schlenker, and J. Costa-Roberts, 2011a: Climate Trends and Global Crop Production Since 1980. Science, 333(6042), 616-620. Lobell, D.B., M. Bänziger, C. Magorokosho, and B. Vivek, 2011b: Nonlinear heat effects on African maize as evidenced by historical yield trials. 1(1), 42-45. Subject to Final Copyedit 49 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Lockwood, A.H., 2012: The Silent Epidemic: Coal and the Hidden Threat to Health. MIT Press, , pp. 248. Lowe, D., K.L. Ebi, and B. Forsberg, 2011: Heatwave early warning systems and adaptation advice to reduce human health consequences of heatwaves. Int J Environ Res Public Health, 8(12), 4623-4648. 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.D. Monache, A.J. Monaghan, D.F. Steinhoff, and L. Eisen, 2012: The Dengue Virus Mosquito Vector Aedes aegypti at High Elevation in M√©xico. The American Journal of Tropical Medicine and Hygiene, 87(5), 902-909. Lu, L. and H. Lin, 2009: Time series analysis of dengue fever and weather in Guangzhou, China. BMC Public Health, 9(395). Luber, G. and M. McGeehin, 2008: Climate Change and Extreme Heat Events. American Journal of Preventive Medicine, 35(5), 429-435. Lucas, R.M., P. Valery, I. van der Mei, T. Dwyer, M.P. Pender, B. Taylor, A. Ponsonby, and G. Ausimmune Investigator, 2013: Sun exposure over a lifetime in Australian adults from latitudinally diverse regions. Photochemistry & Photobiology, 89(3), 737-44. Luginbuhl, R., L. Jackson, D. Castillo, and K. Loringer, 2008: Heat-related deaths among crop workers - United States, 1992 - 2006. JAMA: The Journal of the American Medical Association, 300(9), 1017-1018. Lukan, M., E. Bullova, and B. Petko, 2010: Climate warming and tick-borne encephalitis, Slovakia. Emerg Infect Dis, 16(3), 524-6. Lunde, T.M., M.N. Bayoh, and B. Lindtjorn, 2013: How malaria models relate temperature to malaria transmission. Parasites & Vectors, 6, 20. Lutsey, P.L., M.A. Pereira, A.G. Bertoni, N.R. Kandula, and D.R. Jacobs, 2010: Interactions Between Race/Ethnicity and Anthropometry in Risk of Incident Diabetes. American Journal of Epidemiology, 172(2), 197-204. Lutz, W., W. Sanderson, and S. Scherbov, 2008: The coming acceleration of global population ageing. Nature, 451(7179), 716-719. 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(1), 226. Maas, J., R.A. Verheij, S. de Vries, P. Spreeuwenberg, F.G. Schellevis, and P.P. Groenewegen, 2009: Morbidity is related to a green living environment. J Epidemiol Community Health, 63(12), 967-73. Macdonald, W.W., 1956: Aedes aegypti in Malaysia. II. Larval and adult biology. Annals of Tropical Medicine and Parasitology, 50(4), 399-414. MacMillan, K., A.J. Monaghan, T. Apangu, K.S. Griffith, P.S. Mead, S. Acayo, R. Acidri, S.M. Moore, J.T. Mpanga, R.E. Enscore, K.L. Gage, and R.J. Eisen, 2012: Climate Predictors of the Spatial Distribution of Human Plague Cases in the West Nile Region of Uganda. The American Journal of Tropical Medicine and Hygiene, 86(3), 514-523. Malik, S.M., H. Awan, and N. Khan, 2012: Mapping vulnerability to climate change and its repercussions on human health in Pakistan. Global Health, 8, 31. Mallick, D.L., A. Rahman, M. Alam, A.S.M. Juel, A.N. Ahmad, and S.S. Alam, 2005: Case Study 3: Bangladesh Floods in Bangladesh: A Shift from Disaster Management Towards Disaster Preparedness. IDS Bulletin, 36(4), 53-70. Maloney, S. and C. 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. 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), e1438. Markandya, A., B.G. Armstrong, S. Hales, A. Chiabai, P. Criqui, S. Mima, C. Tonne, and P. Wilkinson, 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: low-carbon electricity generation. The 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. Martin, D., B. Belanger, P. Gosselin, J. Brazeau, C. Furgal, and S. Dery, 2007: Drinking water and potential threats to human health in Nunavik: Adaptation strategies under climate change conditions. Arctic, 60(2), 195-202. McCormack, G. and A. Shiell, 2011: In search of causality: a systematic review of the relationship between the built environment and physical activity among adults. International Journal of Behavioral Nutrition and Physical Activity, 8(1), 125. Subject to Final Copyedit 50 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 McCracken, J.P., K.R. Smith, A. Diaz, M.A. Mittleman, and J. Schwartz, 2007: Chimney stove intervention to reduce long-term wood smoke exposure lowers blood pressure among Guatemalan women. Environ Health Perspect, 115(7), 996-1001. McDonald, R.I., P. Green, D. Balk, B.M. Fekete, C. Revenga, M. Todd, and M. Montgomery, 2011: Urban growth, climate change and freshwater availability. Pnas, 108(15), 6312-6317. McMichael, A.J., J.W. Powles, C.D. Butler, and R. Uauy, 2007: Food, livestock production, energy, climate change, and health. Lancet, 370(9594), 1253-63. McMichael, A.J., P. Wilkinson, R.S. Kovats, S. Pattenden, S. Hajat, B. Armstrong, N. Vajanapoom, E.M. Niciu, H. Mahomed, C. Kingkeow, M. Kosnik, M.S. O'Neill, I. Romieu, M. Ramirez-Aguilar, M.L. Barreto, N. Gouveia, and B. Nikiforov, 2008: International study of temperature, heat and urban mortality: the ISOTHURM project. International Journal of Epidemiology, 37(5), 1121-1131. McMichael, A.J., 2013a: Human Health. In: Four Degrees of Global Warming: Australia in a Hot World. [Christoff, P. (ed.)]. Routledge, London. McMichael, A.J., 2013b: Globalization, climate change, and human health. New England Journal of Medicine, 368(14), 1335-1343. McMichael, C., J. Barnett, and A.J. McMichael, 2012: An Ill Wind? Climate Change, Migration, and Health. Environmental Health Perspectives, 120(5), 646-654. Medlock, J.M. and A. Vaux, 2011: Assessing the possible implications of wetland expansion and management on mosquitoes in Britain. European Mosquito Bulletin, 29, 38-65. Meister, K., C. Johansson, and B. Forsberg, 2012: Estimated short-term effects of coarse particles on daily mortality in Stockholm, Sweden. Environ Health Perspect, 120(3), 431-6. Michon, P., J.L. Cole-Tabian, and E.e.a. Dabod, 2007: The risk of malarial infections and disease in Papua New Guinean children. Am J Trop Med Hyg, 76(6), 997-1008. Milojevic, A., B. Armstrong, S. Kovats, B. Butler, E. Hayes, G. Leonardi, V. Murray, and P. Wilkinson, 2011: Long-term effects of flooding on mortality in England and Wales, 1994-2005: controlled interrupted time-series analysis. Environ Health, 10(11), 1-9. Milojevic, A., B. Armstrong, M. Hashizume, K. McAllister, A. Faruque, M. Yunus, P. Kim Streatfield, K. Moji, and P. Wilkinson, 2012: Health Effects of Flooding in Rural Bangladesh. Epidemiology, 23(1), 107-115. Mitchell, R. and F. Popham, 2007: Greenspace, urbanity and health: relationships in England. J Epidemiol Community Health, 61(8), 681-3. Moore, S.K., N.J. Mantua, and E.P. Salathe Jr., 2011: Past trends and future scenarios for environmental conditions favoring the accumulation of paralytic shellfish toxins in Puget Sound shellfish. Harmful Algae, 10(5), 521-529. Morabito, M., F. Profili, A. Crisci, P. Francesconi, G. Gensini, and S. Orlandini, 2012: Heat-related mortality in the Florentine area (Italy) before and after the exceptional 2003 heat wave in Europe: an improved public health response? International Journal of Biometeorology, 56(5), 801-810. Morello-Frosch, R., M. Zuk, M. Jerrett, B. Shamasunder, and A.D. Kyle, 2011: Understanding the cumulative impacts of inequalities in environmental health: implications for policy. Health Aff (Millwood), 30(5), 879-87. Murray, C.J.L., L.C. Rosenfeld, S.S. Lim, K.G. Andrews, K.J. Foreman, D. Haring, N. Fullman, M. Naghavi, R. Lozano, and A.D. Lopez, 2012: Global malaria mortality between 1980 and 2010: a systematic analysis. The Lancet, 379, 413-431. Nakazawa, Y., R. Williams, A.T. Peterson, P. Mead, E. Staples, and K.L. Gage, 2007: Climate change effects on plague and tularemia in the United States. Vector Borne Zoonotic Dis, 7(4), 529-40. Neelormi, S., N. Adri, and A. Ahmed, 2009: Gender dimensions of differential health effects of climate change induced water-logging: A case study from coastal Bangladesh. Proceedings of IOP Conference Series: Earth and Environmental Science, 2009, Vol 6. 2009. Nelson, G.C., M.W. Rosegrant, A. Palazzo, I. Gray, C. Ingersoll, R. Robertson, S. Tokgov, T. Zhu, S. T.B., C. Ringler, S. Msangi, and L. You, 2010: Food security, farming and climate change to 2050. Scenarios, results, policy options. International Food Policy Research Institute, Washington, D.C. 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 (IFPRI). Nemet, G.F., T. Holloway, and P. Meier, 2010: Implications of incorporating air-quality co-benefits into climate change policymaking. Environmental Research Letters, 5(1), 1-9. Subject to Final Copyedit 51 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Neria, Y., 2012: Mental Health Effects of Hurricane SandyCharacteristics, Potential Aftermath, and Response. JAMA: The Journal of the American Medical Association, 308(24), 2571. Nitschke, M., G. Tucker, A. Hansen, S. Williams, Y. Zhang, and B. Peng, 2011: Impact of two recent extreme heat episodes on morbidity and mortality in Adelaide, South Australia: a case-series analysis. Environmental Health, 10(1), 42-51. Nunn, P.D., 2009: Responding to the challenges of climate change in the Pacific Islands: management and technological imperatives. Climate Research, 40, 211-231. O'Brien, K., M. Pelling, A. Patwardhan, S. Hallegatte, A. Maskrey, T. Oki, U. Oswald-Spring, T. Wilbanks, and P.Z. Yanda, 2012: Toward a sustainable and resilient future. 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. 437-486. OECD, 2012: OECD Environmental Outlook to 2050: Consequences of Inaction. OECD Publishing. Ogden, N.H., A. Maarouf, I.K. Barker, M. Bigras-Poulin, L.R. Lindsay, M.G. Morshed, C.J. O'callaghan, F. Ramay, D. Waltner-Toews, and D.F. Charron, 2006: Climate change and the potential for range expansion of the Lyme disease vector Ixodes scapularis in Canada. International Journal for Parasitology, 36(1), 63-70. 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. 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. Int J Health Geogr, 7, 24. Ogden, N.H., C. Bounchard, K. Kurtenbach, G. Margos, L.R. Lindsay, L. Trudel, S. Nguon, and F. Milord, 2010: Active and passive surveillance and phylogenetic analysis of Borrelia burgdorferi elucidate the process of Lyme disease risk emergence in Canada. Environmental Health Perspectives, 118(7), 909-914. Oikonomou, E. and P. Wilkinson, 2012: Modelling the relative importance of the urban heat island and the thermal quality of dwellings for overheating in London. Building and Environment, 57, 223-238. Okuthe, O.S. and G.E. Buyu, 2006: Prevalence and incidence of tick-borne diseases in smallholder farming systems in the western-Kenya highlands. Vet Parasitol, 141(3-4), 307-12. Omumbo, J., B. Lyon, S. Waweru, S. Connor, and M. Thomson, 2011: Raised temperatures over the Kericho tea estates: revisiting the climate in the East African highlands malaria debate. Malaria Journal, 10(12). O'Neill, M.S., D.K. Jackman, M. Wyman, X. Manarolla, C.J. Gronlund, D.G. Brown, S.J. Brines, J. Schwartz, and A.V. Diez-Roux, 2010: US local action on heat and health: are we prepared for climate change? International Journal of Public Health, 55(2), 105-12. Ostro, B., S. Rauch, R. Green, B. Malig, and R. Basu, 2010: The Effects of Temperature and Use of Air Conditioning on Hospitalizations. American Journal of Epidemiology, 172(9), 1053-1061. 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, 2010: Influence of climate on malaria transmission depends on daily temperature variation. Proc Natl Acad Sci U S A, 107(34), 15135-9. Padmanabha, H., E. Soto, M. Mosquera, C.C. Lord, and L.P. Lounibos, 2010: Ecological links between water storage behaviors and Aedes aegypti production: implications for dengue vector control in variable climates. Ecohealth, 7(1), 78-90. Paerl, H.W., N.S. Hall, and E.S. Calandrino, 2011: Controlling harmful cyanobacterial blooms in a world experiencing anthropogenic and climatic-induced change. Science of the Total Environment, 409(10), 1739- 1745. Page, L.A., S. Hajat, and R.S. Kovats, 2007: Relationship between daily suicide counts and temperature in England and Wales. The British Journal of Psychiatry : The Journal of Mental Science, 191, 106-112. Palecki, M.A., S.A. Changnon, and K.E. Kunke, 2001: The nature and impacts of the July 1999 heat wave in the midwestern United States: learning from the lessons of 1995. Bull Am Meteorol Soc, 82(7), 1353-1367. Pan, A., Q. Sun, A.M. Bernstein, M.B. Schulze, J.E. Manson, M.J. Stampfer, W.C. Willett, and F.B. Hu, 2012: Red Meat Consumption and Mortality: Results From 2 Prospective Cohort Studies. Archives of Internal Medicine, 172(7), 555-563. Pandey, K., 2010: Costs of adapting to climate change for human health in developing countries. In: Development and Climate Change. The World Bank, Washington, DC, pp. 1-19. Subject to Final Copyedit 52 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Paranjothy, S., J. Gallacher, R. Amlot, G.J. Rubin, L. Page, T. Baxter, J. Wight, D. Kirrage, R. McNaught, and P. SR, 2011: Psychosocial impact of the summer 2007 floods in England. BMC Public Health, 11(1), 145. Parsons, K.C., 2003: Human thermal environments: the effects of hot, moderate, and cold environments on human health, comfort, and performance. Taylor & Francis, London ; New York, 2nd ed., pp. 527. Pascual, M., J.A. Ahumada, L.F. Chaves, X. Rodo, and M. Bouma, 2006: Malaria resurgence in the East African highlands: Temperature trends revisited. Pnas, 103(15), 5829-5834. Patt, A.G., M. Tadross, P. Nussbaumer, K. Asante, M. Metzger, J. Rafael, A. Goujon, and G. Brundrit, 2010: Estimating least-developed countries' vulnerability to climate-related extreme events over the next 50 years. Proceedings of the National Academy of Sciences, 107(4), 1333-1337. Patz, J.A., 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. Am J Prev Med, 35(5), 451-8. Paz, S., 2009: Impact of temperature variability on cholera incidence in Southeastern Africa, 1971–2006. Ecohealth, 6(3), 340. Pearce, T., J. Ford, F. Duerden, B. Smit, M. Andrachuk, L. Berrang-Ford, and T. Smith, 2011: Advancing adaptation planning for climate change in the Inuvialuit Settlement Region (ISR): a review and critique. Regional Environmental Change, 11(1), 1-17. Peduzzi, P., B. Chatenoux, H. Dao, A. De Bono, C. Herold, J. Kossin, F. Mouton, and O. Nordbeck, 2012: Global trends in tropical cyclone risk. Nature Climate Change, 2(4), 289-294. Peng, R.D., J.F. Bobb, C. Tebaldi, L. McDaniel, M.L. Bell, and F. Dominici, 2011: Toward a quantitative estimate of future heat wave mortality under global climate change. Environmental Health Perspectives, 119(5), 701-6. Perera, F.P., 2008: Children Are Likely to Suffer Most from Our Fossil Fuel Addiction. Environ Health Perspect, 116(8), 987-990. Pettersson, L., J. Boman, P. Juto, M. Evander, and C. Ahlm, 2008: Outbreak of Puumala virus infection, Sweden. Emerging Infectious Diseases, 14(5), 808-810. Pham, H.V., H.T. Doan, T.T. Phan, and N.N. Minh, 2011: Ecological factors associated with dengue fever in a Central Highlands province, Vietnam. BMC Infect Dis, 11(172), 1-6. Piesse, J. and C. Thirtle, 2009: Three bubbles and a panic: an explanatory review of recent food commodity price events. Food Policy, 34, 119-129. Pitzer, V.E., C. Viboud, L. Simonsen, C. Steiner, C.A. Panozzo, W.J. Alonso, M.A. Miller, R.I. Glass, J.W. Glasser, U.D. Parashar, and B.T. Grenfell, 2009: Demographic variability, vaccination, and the spatiotemporal dynamics of rotavirus epidemics. Science, 325(5938), 290-294. Pitzer, V.E., C. Viboud, B.A. Lopman, M.M. Patel, U.D. Parashar, and B.T. Grenfell, 2011: Influence of birth rates and transmission rates on the global seasonality of rotavirus incidence. Journal of the Royal Society, Interface / the Royal Society, 8(64), 1584-1593. Po, J.Y., J.M. FitzGerald, and C. Carlsten, 2011: Respiratory disease associated with solid biomass fuel exposure in rural women and children: systematic review and meta-analysis. Thorax, 66(3), 232-9. Pokhrel, A.K., M.N. Bates, S.C. Verma, H.S. Joshi, C.T. Sreeramareddy, and K.R. Smith, 2010: Tuberculosis and Indoor Biomass and Kerosene Use in Nepal: A Case–Control Study. Environ Health Perspect, 118(4), 558-564. Polvani, L.M., D.W. Waugh, G.J.P. Correa, and S.W. Son, 2011: Stratospheric Ozone Depletion: The Main Driver of Twentieth-Century Atmospheric Circulation Changes in the Southern Hemisphere. Journal of Climate, 24(3), 795-812. Porter, J.R. and M.A. Semenov, 2005: Crop responses to climatic variation. Philos Trans R Soc Lond B Biol Sci, 360(1463), 2021-2035. Porter, J.R. and M. Gawith, 1999: Temperatures and the growth and development of wheat: a review. European Journal of Agronomy, 10(1), 23-36. Potts, M. and C.E. Henderson, 2012: Global warming and reproductive health. International Journal of Gynecology & Obstetrics, 119, 564-567. Prata, N., 2009: Making family planning accessible in resource-poor settings. Philos Trans R Soc Lond B Biol Sci, 364(1532), 3093-9. PriceWaterhouseCoopers, 2012: Too late for two degrees? Low carbon economy index 2012. PriceWaterhouseCoopers, pp. 1-16. Pudpong, N. and S. Hajat, 2011: High temperature effects on out-patient visits and hospital admissions in Chiang Mai, Thailand. Science of the Total Environment, 409(24), 5260-5267. Subject to Final Copyedit 53 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Puppim de Oliveira, J.A., 2009: The implementation of climate change related policies at the subnational level: An analysis of three countries. Habitat International, 33(3), 253-259. Raffel, T.R., J.M. Romansic, N.T. Halstead, T.A. McMahon, M.D. Venesky, and J.R. Rohr, 2012: Disease and thermal acclimation in a more variable and unpredictable climate. Nature Climate Change, 3(2), 146-151. Ramanathan, V. and G. Carmichael, 2008: Global and regional climate changes due to black carbon. Nature Geoscience, 1(4), 221-227. Ramin, B.M. and A.J. McMichael, 2009: Climate change and health in sub-Saharan Africa: a case-based perspective. Ecohealth, 6(1), 52-7. Ramsey, J.D. and T.E. Bernard, 2000: Heat stress. In: Patty´s Industrial Hygiene, Fifth edition. [Harris, R.L. (ed.)]. Jon Wiley and Sons, New York, pp. 925-985. Ramsey, J., 1995: Task performance in heat: a review. Ergonomics, 38(1), 154-165. Randolph, S.E., 2010: To what extent has climate change contributed to the recent epidemiology of tick-borne diseases? Veterinary Parasitology, 167(2-4), 92-94. Ranjit, S. and N. Kissoon, 2011: Dengue hemorrhagic fever and shock syndromes. Pediatric Critical Care Medicine, 12(1), 90-100. Reed, J. and B. Ainsworth, 2007: Perceptions of environmental supports on the physical activity behaviors of university men and women: a preliminary investigation. J Am Coll Health, 56(2), 199-204. Reid, C.E., M.S. O'Neill, C.J. Gronlund, S.J. Brines, D.G. Brown, A.V. Diez-Roux, and J. Schwartz, 2009: Mapping community determinants of heat vulnerability. Environmental Health Perspectives, 117(11), 1730-6. Ren, C., G.M. Williams, L. Morawska, K. Mengersen, and S. Tong, 2008: Ozone modifies associations between temperature and cardiovascular mortality: analysis of the NMMAPS data. Occup Environ Med, 65(4), 255-60. 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. The American Journal of Tropical Medicine and Hygiene, 84(6), 862-869. Rinaldo, A., E. Bertuzzo, L. Mari, L. Righetto, M. Blokesch, M. Gatto, R. Casagrandi, M. Murray, S.M. Vesenbeckh, and I. Rodriguez-Iturbe, 2012: Reassessment of the 2010–2011 Haiti cholera outbreak and rainfall-driven multiseason projections. Proceedings of the National Academy of Sciences, 109(17), 6602-6607. Rive, N. and K. Aunan, 2010: Quantifying the Air Quality Cobenefits of the Clean Development Mechanism in China. Environmental Science & Technology, 44(11), 4368-4375. Rocklov, J., B. Forsberg, and K. Meister, 2009: Winter mortality modifies the heat-mortality association the following summer. The European Respiratory Journal: Official Journal of the European Society for Clinical Respiratory Physiology, 33(2), 245-51. Rocklov, J. and B. Forsberg, 2009: Comparing approaches for studying the effects of climate extremes - a case study of hospital admissions in Sweden during an extremely warm summer. Global Health Action, 2, 1-11. Rocklov, J., K. Ebi, and B. Forsberg, 2011: Mortality related to temperature and persistent extreme temperatures: a study of cause-specific and age-stratified mortality. Occupational and Environmental Medicine, 68(7), 531-536. Rogelj, J., B. Hare, J. Nabel, K. Macey, M. Schaeffer, K. Markmann, and M. Meinshausen, 2009: Halfway to Copenhagen, no way to 2 °C. Nature Reports Climate Change, 3(July), 1-3. Ronan, K.R., K. Crellin, D.M. Johnston, K. Finnis, D. Paton, and J. Becker, 2008: Promoting Child and Family Resilience to Disasters: Effects, Interventions and Prevention Effectiveness. Children, Youth and Environments, 18(1), 332-353. Rundle, A., K.M. Neckerman, L. Freeman, G.S. Lovasi, M. Purciel, J. Quinn, C. Richards, N. Sircar, and C. Weiss, 2009: Neighborhood food environment and walkability predict obesity in New York City. Environ Health Perspect, 117(3), 442-7. Russell, R.C., B.J. Currie, M.D. Lindsay, J.S. Mackenzie, S.A. Ritchie, and P.I. Whelan, 2009: Dengue and climate change in Australia: predictions for the future should incorporate knowledge from the past. Med J Aust, 190(5), 265-8. Rutstein, S.O., 2005: Effects of preceding birth intervals on neonatal, infant and under-five years mortality and nutritional status in developing countries: evidence from the demographic and health surveys. Int J Gynaecol Obstet, 89 Suppl 1, S7-24. Sahu, S., M. Sett, and T. Kjellstrom, 2013: Heat exposure, cardiovascular stress and work productivity in rice harvesters in India: implications for a climate change future. Industrial Health, 51(4), 424-431. Subject to Final Copyedit 54 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Salomon, J.A., H. Wang, M.K. Freeman, T. Vos, A.D. Flaxman, A.D. Lopez, and C.J. Murray, 2012: Healthy life expectancy for 187 countries, 1990-2010: a systematic analysis for the Global Burden Disease Study 2010. The Lancet, 380(9859), 2144-2162. Samson, J., D. Berteaux, B.J. McGill, and M.M. Humphries, 2011: Geographic disparities and moral hazards in the predicted impacts of climate change on human populations. Global Ecology and Biogeography, 20(4), 532-544. Sankoh, O. and P. Byass, 2012: The INDEPTH Network: filling vital gaps in global epidemiology. International Journal of Epidemiology, 41(3), 579-588. Satish, U., M.J. Mendell, K. Shekhar, T. Hotchi, D. Sullivan, S. Streufert, and W.J. Fisk, 2012: Is CO2 an Indoor Pollutant? Direct Effects of Low-to-Moderate CO2 Concentrations on Human Decision-Making Performance. Environmental Health Perspectives, 120(12), 1671-1677. Sauerborn, R. and K. Ebi, 2012: Climate change and natural disasters – integrating science and practice to protect health. Glob Health Action, 5, 1-7. Schlenker, W. and M.J. Roberts, 2009: Nonlinear temperature effects indicate severe damages to U.S. crop yields under climate change. Pnas, 106(37), 15594-15598. Schmier, J.K. and K.L. Ebi, 2009: The impact of climate change and aeroallergens on children's health. Allergy Asthma Proc, 30(3), 229-37. Schnitzler, J., J. Benzler, D. Altmann, I. Muecke, and G. Krause, 2007: Survey on the population's needs and the public health response during floods in Germany 2002. J Public Health Manag Pract, 13(5), 461-464. Schulte, P.A. and H. Chun, 2009: Climate change and occupational safety and health: establishing a preliminary framework. Journal of Occupational & Environmental Hygiene, 6(9), 542-54. Selin, N.E., S. Wu, K.M. Nam, J.M. Reilly, S. Paltsev, R.G. Prinn, and M.D. Webster, 2009: Global health and economic impacts of future ozone pollution. Environmental Research Letters, 4(044014), 1-9. Sheffield, P.E. and P.J. Landrigan, 2011: Global climate change and children's health: threats and strategies for prevention. Environ Health Perspect, 119(3), 291-8. Sherry, R.A., X. Zhou, S. Gu, J.A. Arnone, D.S. Schimel, P.S. Verburg, L.L. Wallace, and Y. Luo, 2007: Divergence of reproductive phenology under climate warming. Proceedings of the National Academy of Sciences, 104(1), 198-202. Sherwood, S.C. and M. Huber, 2010: An adaptability limit to climate change due to heat stress. Pnas, 107(21), 9552-9555. Shindell, D., J.C.I. Kuylenstierna, E. Vignati, R. van Dingenen, M. Amann, Z. Klimont, S.C. Anenberg, N. Muller, G. Janssens-Maenhout, F. Raes, J. Schwartz, G. Faluvegi, L. Pozzoli, K. Kupiainen, L. Höglund-Isaksson, L. Emberson, D. Streets, V. Ramanathan, K. Hicks, N.T.K. Oanh, G. Milly, M. Williams, V. Demkine, and D. Fowler, 2012: Simultaneously Mitigating Near-Term Climate Change and Improving Human Health and Food Security. Science, 335(6065), 183-189. Shonkoff, S., R. Morello-Frosch, M. Pastor, and J. Sadd, 2011: Environmental Health and Equity Implications of Climate Change and Mitigation Policies in California: A Review of the Literature. 109(S1), S485-503. Silva, H.R., P.E. Phelan, and J.S. Golden, 2010: Modeling effects of urban heat island mitigation strategies on heat- related morbidity: a case study for Phoenix, Arizona, USA. Int J Biometeorol, 54(1), 13-22. Sinha, R., A.J. Cross, B.I. Graubard, M.F. Leitzmann, and A. Schatzkin, 2009: Meat intake and mortality: a prospective study of over half a million people. Arch Intern Med, 169(6), 562-71. Smith, K.R. and K. Balakrishnan, 2009: Mitigating climate, meeting MDGs, and moderating chronic disease: the health co-benefits landscape. In: Commonwealth Health Ministers’ Update 2009. [Commonwealth Secretariat (ed.)]. Commonwealth Secretariat, London, pp. 59-65. Smith, K.R., M. Jerrett, H.R. Anderson, R.T. Burnett, V. Stone, R. Derwent, R.W. Atkinson, A. Cohen, S.B. Shonkoff, D. Krewski, C.A. Pope 3rd, M.J. Thun, and G. Thurston, 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: health implications of short-lived greenhouse pollutants. Lancet, 374(9707), 2091-103. Smith, K.R., 2008: Comparative environmental health assessments. Annals of the New York Academy of Sciences, 1140, 31-9. Smith, K.R. and E. Haigler, 2008: Co-benefits of climate mitigation and health protection in energy systems: scoping methods. Annual Review of Public Health, 29, 11-25. Smith, K.R., J.P. McCracken, M.W. Weber, A. Hubbard, A. Jenny, L.M. Thompson, J. Balmes, A. Diaz, B. Arana, and N. Bruce, 2011: Effect of reduction in household air pollution on childhood pneumonia in Guatemala (RESPIRE): a randomised controlled trial. The Lancet, 378(9804), 1717-1726. Subject to Final Copyedit 55 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Smith, K.R., H. Frumkin, K. Balakrishnan, C.D. Butler, Z.A. Chafe, I. Fairlie, P. Kinney, T. Kjellstrom, D.L. Mauzerall, T.E. McKone, A.J. McMichael, and M. Schneider, 2013: Energy and human health. Annual Review of Public Health, 34, 159-88. Sokolnicki, L.A., N.A. Strom, S.K. Roberts, S.A. Kingsley-Berg, A. Basu, and N. Charkoudian, 2009: Skin blood flow and nitric oxide during body heating in type 2 diabetes mellitus. Journal of Applied Physiology, 106(2), 566-570. Sousa, C.A., M. Clairouin, G. Seixas, B. Viveiros, M.T. Novo, A.C. Silva, M.T. Escoval, and A. Economopoulou, 2012: Ongoing outbreak of dengue type 1 in the Autonomous Region of Madeira, Portugal: preliminary report. Euro Surveill, 17(49), 1-4. Stafoggia, M., F. Forastiere, P. Michelozzi, and C.A. Perucci, 2009: Summer Temperature-related Mortality: Effect Modification by Previous Winter Mortality. Epidemiology, 20(4), 575-583. Stanke, C., M. Kerac, C. Prudhomme, J. Medlock, and V. Murray, 2013: Health effects of drought: a systematic review of the evidence. PLOS Current Disasters, 1(June 5), 1-38. State Environmental Institution “Mosecomonitoring”, 2010: Air quality monitoring in Moscow. WHO Collaborating Centre for Air Quality Management and Air Pollution Control: Newsletter, (46), 9-15. Stenseth, N.C., N.I. Samia, H. Viljugrein, K.L. Kausrud, M. Begon, S. Davis, H. Leirs, V.M. Dubyanskiy, J. Esper, V.S. Ageyev, N.L. Klassovskiy, S.B. Pole, and K.S. Chan, 2006: Plague dynamics are driven by climate variation. Proc Natl Acad Sci U S A, 103(35), 13110-5. 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), e24524. Stone, B., J.J. Hess, and H. Frumkin, 2010: Urban Form and Extreme Heat Events: Are Sprawling Cities More Vulnerable to Climate Change Than Compact Cities? 118, 1425-1428. Strand, L.B., A.G. Barnett, and S. Tong, 2012: Maternal exposure to ambient temperature and the risks of preterm birth and stillbirth in Brisbane, Australia. American Journal of Epidemiology, 175(2), 99-107. Suarez-Varela, M., L. Garci-a-Marcos Alvarez, M. Kogan, A. Gonzalez, A. Gimeno, I. Ontoso, C. Di-az, A. Pena, B. Aurrecoechea, R. Monge, A. Quiros, J. Garrido, I. Canflanca, and A. Varela, 2008: Climate and prevalence of atopic eczema in 6- to 7-year-old school children in Spain. ISAAC PhASE III. International Journal of Biometeorology, 52(8), 833-840. Sudre, B., M. Rossi, W. Van Bortel, K. Danis, A. Baka, and N. Vakalis, 2013: Mapping environmental suitability for malaria transmission, Greece. Emerg Infect Dis, 19(5). Sumilo, D., A. Bormane, L. Asokliene, V. Vasilenko, I. Golovljova, T. Avsic-Zupanc, Z. Hubalek, and S.E. Randolph, 2008: Socio-economic factors in the differential upsurge of tick-borne encephalitis in Central and Eastern Europe. Reviews in Medical Virology, 18(2), 81-95. Tagaris, E., K.J. 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. Environ Sci Technol, 43(13), 4979-88. 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. Int J Biometeorol, 51(3), 193-200. Tate, J.E., C.A. Panozzo, D.C. Payne, M.M. Patel, M.M. Cortese, A.L. Fowkes, and U.D. Parashar, 2009: Decline and changes in seasonality of US rotavirus activity after the introduction of a rotavirus vaccine. Pediatrics, 124(2), 465-471. Tate, J.E., A.H. Burton, C. Boschi-Pinto, A.D. Steele, J. Duque, U.D. Parashar, and WHO-coordinated Global Rotavirus Surveillance Network, 2012: 2008 Estimate of Worldwide Rotavirus-Associated Mortality in Children Younger than 5 Years before the Introduction of Universal Rotavirus Vaccination Programmes: a Systematic Review and Meta-Analysis. The Lancet Infectious Diseases, 12(2), 136-141. Teague, B., R. McLeod, and S. Pascoe, 2010: 2009 Victorian Bushfires Royal Commission, Final Report. Volume 1. Parliament of Victoria, 2009 Victorian Bushfires Royal Commission, Melbourne, pp. 1-42. Teixeira, E.I., G. Fischer, and H. van Velthuizen, 2011: Global hot-spots of heat stress on agricultural crops due to climate change. Agricultural and Forest Meteorology, 170, 206-215. Thompson, A.A., L. Matamale, and S.D. Kharidza, 2012: Impact of Climate Change on Children’s Health in Limpopo Province, South Africa. Int. J. Environ. Res. Public Health, 9, 831-854. Thomson, M.C., F.J. Doblas-Reyes, S.J. Mason, R. Hagedorn, S.J. Connor, T. Phindela, A.P. Morse, and T.N. Palmer, 2006: Malaria early warnings based on seasonal climate forecasts from multi-model ensembles. Nature, 439(7076), 576-579. Subject to Final Copyedit 56 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Thornton, P., P. Jones, T. Owiyo, R. Kruska, M. Herrero, O. V, S. Bhadwal, P. Kristjanson, A. Notenbaert, and N. Bekele, 2008: Climate change and poverty in Africa: mapping hotspots of vulnerability. African Journal of Agriculture and Resource Economics, 2(1), 24-44. Tilman, D., R. Socolow, J.A. Foley, J. Hill, E. Larson, L. Lynd, S. Pacala, J. Reilly, T. Searchinger, C. Somerville, and R. Williams, 2009: Beneficial Biofuels—The Food, Energy, and Environment Trilemma. Science, 325(5938), 270-271. Tirado, M.C., R. Clarke, L.A. Jaykus, A. McQuatters-Gollop, and J.M. Frank, 2010: Climate change and food safety: A review. Climate Change and Food Science, 43(7), 1745-1765. Tokarevich, N.K., A.A. Tronin, O.V. Blinova, R.V. Buzinov, V.P. Boltenkov, E.D. Yurasova, and J. Nurse, 2011: The impact of climate change on the expansion of Ixodes persulcatus habitat and the incidence of tick-borne encephalitis in the north of European Russia. Glob Health Action, 4, 8448. Tollefsen, P., K. Rypdal, A. Torvanger, and N. Rive, 2009: Air pollution policies in Europe: efficiency gains from integrating climate effects with damage costs to health and crops. Environmental Science & Policy, 12(7), 870- 881. Tsai, D.H., J.L. Wang, C.H. Wang, and C.C. Chan, 2008: A study of ground-level ozone pollution, ozone precursors and subtropical meteorological conditions in central Taiwan. J Environ Monit, 10(1), 109-18. Tsui, A.O., A.A. Creanga, and S. Ahmed, 2007: The role of delayed childbearing in the prevention of obstetric fistulas. Int J Gynaecol Obstet, 99 Suppl 1, S98-107. Turcios, R.M., A.T. Curns, R.C. Holman, I. Pandya-Smith, A. LaMonte, J.S. Bresee, R.I. Glass, and National Respiratory and Enteric Virus Surveillance System Collaborating Laboratories, 2006: Temporal and geographic trends of rotavirus activity in the United States, 1997-2004. The Pediatric Infectious Disease Journal, 25(5), 451-454. Uejio, C.K., O.V. Wilhelmi, J.S. Golden, D.M. Mills, S.P. Gulino, and J.P. Samenow, 2011: Intra-urban societal vulnerability to extreme heat: The role of heat exposure and the built environment, socioeconomics, and neighborhood stability. Geographies of Care, 17(2), 498-507. Ujah, I.A., O.A. Aisien, J.T. Mutihir, D.J. Vanderjagt, R.H. Glew, and V.E. Uguru, 2005: Factors contributing to maternal mortality in north-central Nigeria: a seventeen-year review. Afr J Reprod Health, 9(3), 27-40. UNDP, 2011: Human Development Report 2011. Sustainability and Equity: A better future for all. UNDP, New York, pp. 185. UNEP, 2011: Integrated Assessment of Black Carbon and Tropospheric Ozone. United Nations Energy Program and World Meteorological Organization, Nairobi. UNISDR, 2011: Global assessment report on disaster risk reduction. [United Nations International Strategy for Disaster Reduction (ed.)]. United Nations Office for Disaster Risk Reduction, Geneva, pp. 1-178. United Nations Human Settlements Programme, 2011: Cities and climate change: Global report on human settlements 2011. Earthscan, London, pp. 1-279. US EPA, 2010: Air quality trends. United States Environmental Protection Agency. van den Berg, A.E., J. Maas, R.A. Verheij, and P.P. Groenewegen, 2010: Green space as a buffer between stressful life events and health. Soc Sci Med, 70(8), 1203-10. van der Leun, J.C., R.D. Piacentini, and F.R. de Gruijl, 2008: Climate change and human skin cancer. Photochem Photobiol Sci, 7(6), 730-3. van Dillen, S.M.E., S. de Vries, P.P. Groenewegen, and P. Spreeuwenberg, 2011: Greenspace in urban neighbourhoods and residents' health: adding quality to quantity. Journal of Epidemiology and Community Health, 66(6), e8. Van Kerkhove, M.D., K.A.H. Vandemaele, V. Shinde, G. Jaramillo-Gutierrez, A. Koukounari, C.A. Donnelly, L.O. Carlino, R. Owen, B. Paterson, L. Pelletier, J. Vachon, C. Gonzalez, Y. Hongjie, F. Zijian, K.C. Shuk, A. Au, S. Buda, G. Krause, W. Haas, and I. Bonmarin, 2011: Risk Factors for Severe Outcomes following 2009 Influenza A (H1N1) Infection: A Global Pooled Analysis. PLoS Medicine, 8(7), 1-12. Van Kleef, E., H. Bambrick, and S. Hales, 2010: The geographic distribution of dengue fever and the potential influence of global climate change. [Special Programme for Research and Training in Tropical Diseases (ed.)], pp. 1-22. Vedal, S. and S.J. Dutton, 2006: Wildfire air pollution and daily mortality in a large urban area. Environ Res, 102(1), 29-35. Venkataraman, C., A.D. Sagar, G. Habib, N. Lam, and K.R. Smith, 2010: The Indian National Initiative for Advanced Biomass Cookstoves: The benefits of clean combustion. 14(2), 63-72. Subject to Final Copyedit 57 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Vezzulli, L., I. Brettar, E. Pezzati, P.C. Reid, R.R. Colwell, M.G. Hofle, and C. Pruzzo, 2012: Long-term effects of ocean warming on the prokaryotic community: evidence from the vibrios. Isme j, 6(1), 21-30. Wang, L., 2012: Association between the incidence of typhoid and paratyphoid fever and meteorological variables in Guizhou, China. Chinese Medical Journal, 125(3), 455. Wang, X., R. Chen, X. Meng, F. Geng, C. Wang, and H. Kan, 2013: Associations between fine particle, coarse particle, black carbon and hospital visits in a Chinese city. Science of the Total Environment, 458-460, 1-6. Weis, K.E., 2011: Vibrio illness in Florida, 1998–2007. Epidemiology and Infection, 139(4), 591. Weisent, J., W. Seaver, A. Odoi, and B. Rohrbach, 2010: Comparison of three time-series models for predicting campylobacteriosis risk. Epidemiology and Infection, 138(6), 898-906. Weisskopf, M.G., H.A. Anderson, S. Foldy, L.P. Hanrahan, K. Blair, and T.J. Török, 2002: Heat wave morbidity and mortality, Milwaukee, Wis, 1999 vs 1995: an improved response? Am J Public Health, 29(5), 830-833. West, J.J., A. Fiore, and L. Horowitz, 2012: Scenarios of methane emission reductions to 2030: abatement costs and co-benefits to ozone air quality and human mortality. Climatic Change, 114(3-4), 441-461. West, J.J., A.M. Fiore, V. Naik, L.W. Horowitz, M.D. Schwarzkopf, and D.L. Mauzerall, 2007: Ozone air quality and radiative forcing consequences of changes in ozone precursor emissions. Geophysical Research Letters, 34(6), 1-6. West, J.J., S.J. Smith, R.A. Silva, V. Naik, Y. Zhang, Z. Adelman, M.M. Fry, S. Anenberg, L.W. Horowitz, and J. Lamarque, 2013: Co-benefits of mitigating global greenhouse gas emissions for future air quality and human health. 3(10), 885-889. Wilkinson, P., K.R. Smith, S. Beevers, C. Tonne, and T. Oreszczyn, 2007: Energy, energy efficiency, and the built environment. Lancet, 370(9593), 1175-87. 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. Wilkinson, P., K.R. Smith, M. Joffe, and A. Haines, 2007: A global perspective on energy: health effects and injustices. Lancet, 370(9591), 965-78. Willett, K.M. and S. Sherwood, 2012: Exceedance of heat index thresholds for 15 regions under a warming climate using the wet-bulb globe temperature. International Journal of Climatology, 32(2), 161-177. 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. Woodcock, J., P. Edwards, C. Tonne, B.G. Armstrong, O. Ashiru, D. Banister, S. Beevers, Z. Chalabi, Z. Chowdhury, A. Cohen, O.H. Franco, A. Haines, R. Hickman, G. Lindsay, I. Mittal, D. Mohan, G. Tiwari, A. Woodward, and I. Roberts, 2009: Public health benefits of strategies to reduce greenhouse-gas emissions: urban land transport. Lancet, 374(9705), 1930-43. Woodcock, J., M. Givoni, and A.S. Morgan, 2013: Health Impact Modelling of Active Travel Visions for England and Wales Using an Integrated Transport and Health Impact Modelling Tool (ITHIM). Plos One, 8(1), e51462. Woodward, A., G. Lindsay, and S. Singh, 2011: Adapting to climate change to sustain health. WIREs Clim Change, 2(2), 271-282. World Bank, 2010: World Development Report 2010. Development and Climate Change. The World Bank, Washington, DC, pp. xxi, 417 p. World Health Organization, 2003: WHO Guide to Cost-Effectiveness Analysis. [Tan-Torres Edejer, T., Baltussem, R., Adam, T., Hutubessy, R., Acharya, A., Evans, D.B. and Murray, C.J.L. (eds.)]. World Health Organization, Geneva, pp. 1-329. World Health Organization, 2006: WHO Air quality guidelines for particulate matter, ozone, nitrogen dioxide and sulfur dioxide: Global update 2005. World Health Organization, Geneva, pp. 1-22. World Health Organization, 2008a: The Global Burden of Disease. 2004 Update. WHO, Geneva. World Health Organization, 2008b: World Malaria Report 2008. World Health Organization, Geneva, pp. 1-190. World Health Organization, 2009a: WHO Guide to identifying the economic consequences of disease and injury. World Health Organization, Geneva. World Health Organization, 2009b: Save Lives: Make Hospitals Safe in Emergencies. WHO/HAC/WHD/2009.1. World Health Organization, Geneva. World Health Organization, 2010: World Malaria Report 2010. World Health Organization, Geneva, pp. 1-93. World Health Organization, 2011: Gender, Climate Change and Health. WHO, Geneva. World Health Organization, 2012: World Malaria Report 2012. World Health Organization, Geneva, pp. 1-195. Subject to Final Copyedit 58 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 World Health Organization, 2013: Impact of Dengue World Health Organization, Geneva, pp. 3. World Health Organization and World Meteorological Organization, 2012: Atlas of health and climate. [Castonguay, S. (ed.)]. World Health Organization and World Meteorological Organization, Geneva, pp. 1-68. World Health Organization Regional Office for Europe, 2009: Improving public health responses to extreme weather/heat-waves - EuroHEAT: Technical Summary. World Health Organization Regional Office for Europe, Copenhagen, pp. 1-70. World Health Organization Regional Office for Europe, 2010: WHO guidelines for indoor air quality: selected pollutants. World Health Organization, Copenhagen, Denmark, pp. 454. World Health Organization Regional Office for Europe, 2013: Climate change and health: a tool to estimate health and adaptation costs. World Health Organization, Copenhagen, pp. 1-46. Wu, F., Q. Liu, L. Lu, J. Wang, X. Song, and D. Ren, 2011: Distribution of Aedes albopictus (Diptera: Culicidae) in northwestern China. Vector Borne Zoonotic Dis, 11(8), 1181-6. Wu, P.C., J.G. Lay, H.R. Guo, C.Y. Lin, S.C. Lung, and H.J. Su, 2009: Higher temperature and urbanization affect the spatial patterns of dengue fever transmission in subtropical Taiwan. Sci Total Environ, 407(7), 2224-33. Wyndham, C.H., 1969: Adaptation to heat and cold. Environ Res, 2(5), 442-69. Xu, L., Q. Liu, L.C. Stige, T. Ben Ari, X. Fang, K.S. Chan, S. Wang, N.C. Stenseth, and Z. Zhang, 2011: Nonlinear effect of climate on plague during the third pandemic in China. Proc Natl Acad Sci U S A, 108(25), 10214-9. Xu, X., E. Yu, X. Gao, N. Song, L. Liu, X. Wei, W. Zhang, and C. Fu, 2012: Red and processed meat intake and risk of colorectal adenomas: A meta-analysis of observational studies. International Journal of Cancer, 132(2), 437-448. Yoshida, S., T. Satake, and D.S. Mackill, 1981: High-temperature stress in rice [study conducted at IRRI, Philippines]. In: IRRI Research Paper Series No. 67. International Rice Research Institute, pp. 1-15. Zanobetti, A., M.S. O'Neill, C.J. Gronlund, and J.D. Schwartz, 2012: Summer temperature variability and long-term survival among elderly people with chronic disease. Proceedings of the National Academy of Sciences, 10(17), 6608-6613. Zhang, J. and K. Smith, 2007: Household air pollution from coal and biomass fuels in China: measurements, health impacts, and interventions. Environmental Health Perspectives, 115(6), 848-855. Zhang, Y., P. Bi, and J.E. Hiller, 2010: Climate variations and Salmonella infection in Australian subtropical and tropical regions. Science of the Total Environment, 408(3), 524-530. Zhang, Y., P. Bi, Y. Sun, and J.E. Hiller, 2012: Projected Years Lost due to Disabilities (YLDs) for bacillary dysentery related to increased temperature in temperate and subtropical cities of China. Journal of Environmental Monitoring, 14(2), 510-516. Zhou, X., G. Yang, K. Yang, X. Wang, Q. Hong, L. Sun, J.B. Malone, T.K. Kristensen, N.R. Bergquist, and J. Utzinger, 2008: Potential Impact of Climate Change on Schistosomiasis Transmission in China. American Journal of Tropical Medicine and Hygiene, 78(2), 188-194. Zhu, B.P., 2005: Effect of interpregnancy interval on birth outcomes: findings from three recent US studies. Int J Gynaecol Obstet, 89 Suppl 1, S25-33. Subject to Final Copyedit 59 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 11-1: The association between different climatic drivers and the global prevalence and geographic distribution of selected vector-borne diseases observed over the period 2008-2012. Among the vector borne diseases shown here, only dengue fever was associated with climate variables at both the global and local levels (high confidence), while malaria and hemorrhagic fever with renal syndrome showed a positive association at the local level (high confidence). Table 11-2: Number of under-nourished children less than 5 years of age (in millions) in 2000 and 2050, using the NCAR (National Center for Atmospheric Research) climate model (and the A2 scenario from AR4). Results assume no effect of heat on farmers’ productivity, and no CO2 fertilization benefits. Adapted from Nelson et al. (2009). Scenario South East Asia/ Europe Latin Middle Sub- All Developing Asia Pacific and America and East/ Saharan Countries Central Caribbean North Africa Asia Africa 2000 75.6 23.8 4.1 7.7 3.5 32.7 147.9 2050 No climate 52.3 10.1 2.7 5.0 1.1 41.7 113.3 change Climate change 59.1 14.5 3.7 6.4 2.1 52.2 138.5 Subject to Final Copyedit 60 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Table 11-3: Examples of recent (post AR4) research studies on co-benefits of climate change mitigation and public health policies. For recent estimates of the global and regional burden of disease from the various risk factors involved, see Lim et al. (2012). Co-benefit category Benefits for health Benefits for climate References in chronological order Reduction of co-pollutants from Potentially reduce exposures that are associated with Reduces CAP emissions associated with Bell et al., 2008 household solid fuel disease, chronic and acute respiratory illnesses, lung household solid fuel use including CO2, CO, Smith et al., 2008 combustion [see also WGIII-7, cancer, low birth weight and stillbirths, and possibly black carbon, and CH4 Wilkinson et al., 2009 WGIII-8, WGIII-9, WGIII-10] tuberculosis Lefohn et al., 2010 Venkataraman et al., 2010 World Health Organization Regional Office for Europe, 2010 Po et al., 2011 Anenberg et al., 2012 Reduction of GHGs and Reductions health-damaging co-pollutant emissions Reductions in emissions of CO2, black carbon, Bell et al., 2008 associated co-pollutants from would decrease exposures to outdoor air pollution and CO, CH4,and other CAPs Apsimon et al., 2009 industrial sources, such as power could reduce risks of cardiovascular disease, chronic Jacobson, 2009 plants and landfills by more and acute respiratory illnesses, lung cancer, and Puppim de Oliveira et al., 2009 efficient generation or preterm birth. Smith et al., 2009 substitution of low carbon Tollefsen et al., 2009 alternatives [27.3.7.2] Dennekamp et al., 2010 Jacobson, 2010 Nemet, et al., 2010 Rive and Aunan, 2010 Shonkoff et al., 2011 Shindell et al., 2012 West et al., 2012 West et al., 2013 Energy efficiency. Actual energy Reductions in fuel demand potentially can reduce Reductions in emission of CAPs due to Markandya et al., 2009 reduction may sometimes be less emissions of CAPs associated with fuel combustion decreases in fuel consumption Wilkinson et al., 2009 than anticipated because part of and subsequent exposures to pollutants that are known the efficiency benefit is taken as to be health damaging. more service Subject to Final Copyedit 61 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Increases in active travel and Increased physical activity; reduced obesity; reduced Reductions of CAP emissions associated with Babey et al., 2007 reductions in pollution due to non communicable disease burden, health service costs vehicle transport; Replacing existing vehicles Reed and Ainsworth, 2007 modifications to the built averted; improved mental health; reduced exposure to with lower emission vehicles could reduce air Kaczynski and Henderson, 2008 environment, including better air pollution; increased local access to essential pollution Casagrande et al., 2009 access to public transport and services, including food stores; enhanced safety. Jarrett et al., 2009 higher density of urban Rundle et al., 2009 settlements [See also 24.4, 24.5, Woodcock et al., 2009 24.6, 24.7, 26.8,] Durand et al., 2011 Grabow et al., 2011 McCormack and Shiell, 2011 Jensen et al., 2013 Woodcock et al., 2013 Healthy low GHG emission diets Reduced dietary saturated fat in some populations Reductions in CO2 and CH4 emissions from McMichael et al., 2007 which can have beneficial (particularly from ruminants) and replacement by plant energy-intensive livestock systems Friel et al., 2009 effects on a range of health sources associated with decreased risk of (ischemic) Sinha et al., 2009 outcomes [See also Table 11.3] heart disease, stroke, colorectal cancer (processed meat Smith and Balakrishnan, 2009 consumption) Increased fruit and vegetable Jakszyn et al., 2011 consumption can reduce risk of chronic diseases. Hooper et al., 2012 Reduced CH4 emissions due to a decreased demand for Pan et al., 2012 ruminant meat products would reduce the tropospheric Xu et al., 2012 ozone. Greater access to reproductive Lower child and maternal mortality from increased Potentially slower growth of energy Tsui et al. 2007 health services birth intervals and shifts in maternal age. consumption and related CAP emissions; less Gribble et al., 2009 impact on land use change, etc. Prata, 2009 O'Neill et al., 2010 Diamond-Smith and Potts, 2011 Potts and Henderson, 2012 Kozuki et. al., 2013 Increases in urban green space Reduced temperatures and heat island effects; reduced Reduces atmospheric CO2 via carbon Mitchell and Popham, 2007 [Table 25-5] noise; enhanced safety; psychological benefits; better sequestration in plant tissue and soil Babey et al., 2008 self-perceived health status. Maas et al., 2009 van den Berg et al., 2010 van Dillen et al., 2011 Carbon sequestration forest Poverty alleviation and livelihood/job generation Reduces emissions of CAPs and promotes Holmes, 2010 plantations, REDD and carbon through sale of CDM and voluntary market credits. carbon sequestration through REDD Ezzine-de-Blas et al., 2011 offset sales [see Chpt 13, 15.3.4, Ameliorate declines in production or competitiveness see also 20.4.1, 26.8.4.3] in rural communities. Subject to Final Copyedit 62 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 11-1: Conceptual diagram showing three primary exposure pathways by which climate change affects health: directly through weather variables such as heat and storms; indirectly through natural systems such as disease vectors; and pathways heavily mediated through human systems such as undernutrition. The yellow box indicates the moderating influences of local environmental conditions on how climate change exposure pathways are manifest in a particular population. The orange box indicates that the extent to which the three categories of exposure translate to actual health burden is moderated by such factors as background public health and socioeconomic conditions, and adaptation measures. The green arrows at the bottom indicate that there may be feedback mechanisms, positive or negative, between societal infrastructure, public health, and adaptation measures and climate change itself. As discussed later in the chapter, for example, some measures to improve health also reduce emissions of climate-altering pollutants, thus reducing the extent and/or pace of climate change as well as improving local health. Credit: E. Garcia, UC Berkeley. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 63 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 11-2: Increasingly frequent heat extremes will combine with rapidly growing numbers of older people living in cities – who are particularly vulnerable to extreme heat. Countries are shaded according to the expected proportional increase in urban populations aged over 65 by the year 2050. Bar graphs show how frequently the maximum daily temperature that would have occurred only once in 20 years in the late 20th century is expected to occur in the mid-21st century, with lower numbers indicating more frequent events. Results are shown for 3 different “SRES” scenarios (Blue = B1; Green = A1B, Red = A2), as described in the IPCC Special Report on Emissions Scenarios, and based on 12 global climate models participating in the third phase of the Coupled Model Intercomparison Project (CMIP3). Coloured boxes show the range in which 50% of the model projections are contained, and whiskers show the maximum and minimum projections from all models. Source: World Health Organization and World Meteorological Organization, 2012. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 64 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Rainfall, Temperature, Breteau index and Dengue cases, Trinidad (2002-2004). 1600 50 45 Temperature and Breteau index 1400 No. dengue cases, Rainfall 40 1200 35 1000 30 800 25 600 20 15 400 10 200 5 0 0 n n n v. v. v. ch ly ch ly ch ly . . . ay ay ay pt pt pt Ja Ja Ja Ju Ju Ju No No No ar ar ar M M M Se Se Se M M M Months Rainfall (mm) Dengue cases Temp Breteau Fig. 3. Efficacy of pre-seasonal treatment with temephos on Aedes aegypti ovitrap egg counts in Curepe (treatment) and St. Joseph (control), Trinidad (2003) 2000 1500 No. of Eggs Treatment 1000 Control 500 0 1 3 5 7 9 11 13 15 17 19 21 Weeks Figure 11-3: a) Rainfall, temperature, Breteau index (number of water containers with Ae. aegypti larvae per 100 houses), and dengue fever cases, Trinidad (2002-2004). Rainfall was found to be significantly correlated with an increase in the Ae. aegypti population and dengue fever incidence, with a clearly defined “dengue season” between June and November over two years of the study. Source: (Chadee et al., 2007). b) Efficacy of pre-seasonal treatment with temephos on Ae. aegypti ovitrap egg counts in Curepe (treatment) and St. Joseph (control), Trinidad (2003). Evidence of the efficacy of the pre-seasonal larval control through focal treatment of Ae. aegypti population is provided. Treatment at the onset of the rainy season can effectively prevent the rapid increase in Ae. aegypti populations and therefore suppress the onset of dengue transmission. Source: Chadee, 2009. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 65 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 11-4: Effect of rising temperatures on the area in which transmission of Schistosomiasis japonica may occur. Green area denotes the range of schistosomiasis in China in 2000. The blue area shows the additional area suitable for disease transmission in 2050. Based on a biology-driven model including parasite (Schistosoma japonicum) and snail intermediate host (Oncomelania hupensis) and assuming average temperatures in China in mid-winter (January) increase by 1.6°C in 2050, compared with 2000. Adapted from Zhou et al., 2008. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 66 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Job exertion required (Watts) 500 400 300 200 100 90 e.g. 2oC rise This figure to be inserted in the 80 approx. Pacific Ocean area of the map 70 halves above Percent 60 work of full output 50 working capacity 40 30 20 10 0 20 oC 25 oC 30 oC 35 oC 40 oC Wet Bulb Globe Temperature (WBGT) Figure 11-5: The 1980-2009 average of the hottest months globally, measured in web bulb globe temperature (WBGT), which combines temperature, humidity, and other factors into a single index of the impact on work capacity and threat of heat exhaustion. The insert shows the International Standard Organization standard (1989) for heat stress in the workplace that leads to recommendations for increased rest time per hour to avoid heat exhaustion at different work levels. This is based on studies of healthy young workers and includes a margin of safety. Note that some parts of the world already exceed the level for safe work activity during the hottest month. In general, with climate change, for every 1°C that Tmax goes up, the WBGT goes up by about 0.9°C, leading to more parts of the world being restricted for more of the year, with consequent impacts on productivity, heat exhaustion, and need for air conditioning to protect health. Source: Lemke and Kjellstrom, 2012. [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 67 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 11-6: Conceptual presentation of the health impacts from climate change and the potential for impact reduction through adaptation. Impacts are identified in eight health-related sectors based on assessment of the literature and expert judgments by authors of Chapter 11.The width of the slices indicates in a qualitative way the relative importance in terms of burden of ill-health globally at present and should not be considered completely independent. Impact levels are presented for the near-term era of committed climate change (2030-2040), in which projected levels of global mean temperature increases do not diverge substantially across emissions scenarios. For some sectors, e.g., vector-borne diseases, heat/cold stress, and agricultural production and undernutrition, there may be benefits to health in some areas, but the net impact is expected to be negative. Estimated impacts are also presented for the longer-term era of climate options (2080-2100), for global mean temperature increase of 4°C above preindustrial levels, which could potentially be avoided by vigorous mitigation efforts taken soon. For each timeframe, impact levels are estimated for the current state of adaptation and for a hypothetical highly adapted state, indicated by different colors. Subject to Final Copyedit 68 28 October 2013 FINAL DRAFT IPCC WGII AR5 Chapter 11 Do Not Cite, Quote, or Distribute Prior to Public Release on 31 March 2014 Figure 11-7: Illustrative co-benefits comparison of the health and climate cost-effectiveness of selected household, transport, and power sector interventions (Smith and Haigler, 2008). Area of each circle denotes the total social benefit in international dollars from the combined value of carbon offsets (valued at 10$/tCO2e) and averted DALYs [$7450/DALY, which is representative of valuing each DALY at the average world GDP (PPP) per capita in 2000]. The vertical bar shows the range of the cut offs for cost-effective and very cost-effective health interventions in India and China using the WHO CHOICE criteria (World Health Organization, 2003). This figure evaluates only a small subset of all co-benefits opportunities and thus should not be considered either current or complete. It does illustrate, however, the kind of comparisons that can help distinguish and prioritize options. Note that even with the log-log scaling, there are big differences among them. For other figures comparing the climate and health benefits of co-benefits actions including those in food supply and urban design, see Haines et al. (Haines et al., 2009). See the original reference for details of the calculations in this figure (Smith and Haigler, 2008). [Illustration to be redrawn to conform to IPCC publication specifications.] Subject to Final Copyedit 69 28 October 2013