FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Chapter 12: Central and South America 3 4 Coordinating Lead Authors: Edwin J. Castellanos (Guatemala), Maria Fernanda Lemos (Brazil) 5 6 Lead Authors: Laura Astigarraga (Uruguay), Noemí Chacón (Venezuela), Nicolás Cuvi (Ecuador), 7 Christian Huggel (Switzerland), Liliana Miranda (Peru), Mariana Moncassim Vale (Brazil), Jean Pierre 8 Ometto (Brazil), Pablo L. Peri, (Argentina), Julio C. Postigo (USA/Peru), Laura Ramajo (Chile/Spain), 9 Lisandro Roco (Chile), Matilde Rusticucci (Argentina). 10 11 Contributing Authors: Júlia Alves Menezes (Brazil), Pedro Borges (Venezuela), Jhonattan Bueno 12 (Venezuela), Francisco Cuesta (Ecuador), Fabian Drenkhan (Peru), Alex Guerra (Guatemala), Valeria 13 Guinder (Argentina), Isabel Hagen (Switzerland), Jorgelina Hardoy (Argentina), Stella Hartinger (Peru), 14 Gioconda Herrera (Ecuador), Cecilia Herzog (Brazil), Bárbara Jacob (Chile), Thais Kasecker (Brazil), 15 Andrea Lampis (Colombia/Brazil), Izabella Lentino (Brazil), Luis C. S. Madeira Domingues (Brazil), José 16 Marengo (Brazil), David Montenegro Lapola (Brazil), Ana Rosa Moreno (Mexico), Julia de Niemeyer 17 Caldas (Brazil), Eduardo Pacay (Costa Rica/Guatemala), Roberto Pasten (Chile), Matias Piaggio (Uruguay), 18 Osvaldo Rezende (Brazil), Alfonso J. Rodriguez-Morales (Colombia), Marina Romanello (Argentina/United 19 Kingdom), Sadie J. Ryan (USA/ United Kingdom), Anna Stewart-Ibarra (USA/Ecuador), María Valladares 20 (Chile/Spain) 21 22 Review Editors: Carlos Méndez (Venezuela), Avelino Suarez (Cuba) 23 24 Chapter Scientist: María Valladares (Chile/Spain) 25 26 Date of Draft: 1 October 2021 27 28 Notes: TSU Compiled Version 29 30 31 Table of Contents 32 33 Executive Summary..........................................................................................................................................3 34 12.1 Introduction ..............................................................................................................................................8 35 12.1.1 The Central and South America Region .........................................................................................8 36 12.1.2 Approach and Storyline for the Chapter.......................................................................................10 37 12.2 Summary of the Fifth Assessment Report and Recent IPCC Special Reports.................................11 38 12.3 Hazards, Exposure, Vulnerabilities and Impacts ................................................................................12 39 12.3.1 Central America (CA) Sub-region ................................................................................................12 40 12.3.2 Northwest South America (NWS) Sub-region ...............................................................................16 41 12.3.3 Northern South America (NSA) Sub-region..................................................................................20 42 12.3.4 South America Monsoon (SAM) Sub-region.................................................................................23 43 12.3.5 Northeast South America (NES) Sub-region.................................................................................28 44 12.3.6 Southeast South America (SES) Sub-region..................................................................................30 45 12.3.7 Southwest South America (SWS) Sub-region ................................................................................34 46 12.3.8 Southern South America (SSA) Sub-region...................................................................................39 47 12.4 Key Impacts and Risks...........................................................................................................................44 48 12.5 Adaptation...............................................................................................................................................50 49 12.5.1 Terrestrial and Freshwater Ecosystems and their Services..........................................................50 50 12.5.2 Ocean and Coastal Ecosystems and their Services .....................................................................53 51 12.5.3 Water.............................................................................................................................................58 52 12.5.4 Food, Fibre and other Ecosystem Products..................................................................................63 53 12.5.5 Cities, Settlements and Infrastructure...........................................................................................69 54 12.5.6 Health and Wellbeing....................................................................................................................73 55 12.5.7 Poverty, Livelihood and Sustainable Development ......................................................................78 56 12.5.8 Cross-cutting Issues in the Human Dimension .............................................................................82 57 12.5.9 Adaptation Options to Address Key Risks in CSA ........................................................................88 Do Not Cite, Quote or Distribute 12-1 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 12.5.10 Feasibility Assessment of Adaptation Options.....................................................................91 2 12.6 Case Studies ............................................................................................................................................93 3 12.6.1 Nature-based Solutions in Quito, Ecuador ...................................................................................93 4 12.6.2 Anthropogenic Soils, an Option for Mitigation and Adaptation to Climate Change in Central 5 and South America. Learning from the "Terras Pretas de Índio" in the Amazon ...............95 6 12.6.3 Towards a Metropolitan Water-related Climate Proof Governance (re)configuration? The case 7 of Lima, Perú ........................................................................................................................96 8 12.6.4 Strengthening Water Governance for Adaptation to Climate Change: Managing Scarcity and 9 Excess of Water in the Pacific Coastal area of Guatemala..................................................97 10 12.7 Knowledge Gaps .....................................................................................................................................98 11 12.7.1 Knowledge Gaps in the Subregions ..............................................................................................99 12 12.7.2 Knowledge Gaps by Sector ...........................................................................................................99 13 12.8 Conclusion .............................................................................................................................................103 14 FAQ 12.1: How are inequality and poverty limiting options to adapt to climate change in Central and 15 South America? ....................................................................................................................................106 16 FAQ 12.2: How have urban areas in Central and South America adapted to climate change so far, 17 which further actions should be considered within the next decades and what are the limits of 18 adaptation and sustainability? ............................................................................................................107 19 FAQ 12.3: How do climatic events and conditions affect migration and displacement in Central and 20 South America, will this change due to climate change, and how can communities adapt? .........108 21 FAQ 12.4: How is climate change impacting and expected to impact food production in Central and 22 South America in the next 30 years and what effective adaptation strategies are and can be 23 adopted in the region?..........................................................................................................................109 24 FAQ 12.5: How can Indigenous knowledge and practices contribute to adaptation initiatives in 25 Central and South America? ...............................................................................................................111 26 References......................................................................................................................................................113 27 28 Do Not Cite, Quote or Distribute 12-2 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Executive Summary 2 3 Vulnerability and observed impacts: 4 5 Central and South America are highly exposed, vulnerable and strongly impacted by climate change, 6 a situation amplified by inequality, poverty, population growth and high population density, land use 7 change particularly deforestation with the consequent biodiversity loss, soil degradation, and high 8 dependence of national and local economies on natural resources for production of commodities (high 9 confidence1). Profound economic, ethnic and social inequalities are exacerbated by climate change. High 10 levels of widespread poverty, weak water governance, unequal access to safe water and sanitation services 11 and lack of infrastructure and financing reduce adaptation capacity, increasing and creating new population 12 vulnerabilities (high confidence). {12.1.1, 12.2, 12.3, 12.5.5, 12.5.7, Figure 12.2} 13 14 The Amazon forest, one of the world's largest biodiversity and carbon repositories, is highly 15 vulnerable to drought (high confidence). The Amazon forest was highly impacted by the unprecedented 16 droughts and higher temperatures observed in 1998, 2005, 2010 and 2015/2016 attributed partly to climate 17 change. This resulted in high tree mortality rates and basin-wide reductions in forest productivity, 18 momentarily turning pristine forest areas from a carbon sink into a net source of carbon to the atmosphere 19 (high confidence). Other terrestrial ecosystems in Central and South America have been impacted by climate 20 change, through persistent drought or extreme climatic events. The combined effect of anthropogenic land 21 use change and climate change increases the vulnerabilities of terrestrial ecosystems to extreme climate 22 events and fires (medium confidence). {12.3, 12.4, Figure 12.7, Figure 12.9, Figure 12.10} 23 24 The distribution of terrestrial species has changed in the Andes due to increasing temperature (very 25 high confidence). Species have shifted upslope leading to range contractions for highland species, and range 26 contractions and expansions for lowland species, including crops and vectors of diseases (very high 27 confidence). {12.3.2.4} 28 29 Ocean and coastal ecosystems in the region such as coral reefs, estuaries, salt marshes, mangroves and 30 sandy beaches are highly sensitive and negatively impacted by climate change and derived-hazards 31 (high confidence). Observed impacts include the reduction in coral abundance, density and cover in Central 32 America, Northwest South America and Northeast South America and increasing number of coral bleaching 33 events in Central America and Northeast South America; changes in the plankton community and in ocean 34 and coastal food web structures, loss of vegetated wetlands and changes in macrobenthic communities in 35 Central America, Northwest, Northern, and Southeast South America. {12.3, 12.5.2, Figure 12.8, Figure 36 12.9, Table SM12.3} 37 38 Global warming has caused glacier loss in the Andes from 30% to more than 50% of their area since 39 the 1980s. Glacier retreat, temperature increase and precipitation variability, together with land-use 40 change, have affected ecosystems, water resources, and livelihoods through landslides and flood 41 disasters (very high confidence). In several areas of the Andes, flood and landslide disasters have increased, 42 and water availability and quality and soil erosion have been affected by both climatic and non-climatic 43 factors (high confidence). {12.3.2, 12.3.7, Figure 12.9, Figure 12.13, Table SM12.6} 44 45 The scientific evidence since the IPCC AR5 increased the confidence on the synergy among fire, land 46 use change, particularly deforestation, and climate change, directly impacting human health, 47 ecosystem functioning, forest structure, food security and the livelihoods of resource-dependent 48 communities (medium confidence). Regional increase in temperature, aridity and drought increased the 49 frequency and intensity of fire. On average, people in the region were more exposed to high fire danger 50 between 1 and 26 additional days depending on the subregion for the years 2017-2020 compared to 2001- 51 2004 (high confidence). {12.2, 12.3, Figure 12.9, Figure 12.10, Table 12.5} 1 In this Report, the following summary terms are used to describe the available evidence: limited, medium, or robust; and for the degree of agreement: low, medium, or high. A level of confidence is expressed using five qualifiers: very low, low, medium, high, and very high, and typeset in italics, e.g., medium confidence. For a given evidence and agreement statement, different confidence levels can be assigned, but increasing levels of evidence and degrees of agreement are correlated with increasing confidence. Do Not Cite, Quote or Distribute 12-3 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Changes in timing and magnitude of precipitation and extreme temperatures are impacting 3 agricultural production (high confidence). Since the mid-20th century, increasing mean precipitation has 4 positively impacted agricultural production in Southeast South America, although extremely long dry spells 5 have become more frequent affecting the economies of large cities in southeast Brazil. Inversely, reduced 6 precipitation and altered rainfall at the start and end of the rainy season and during the mid-summer drought 7 is impacting rainfed subsistence farming particularly in the Dry Corridor in Central America and in the 8 tropical Andes compromising food security (high confidence). The crop growth duration for maize for those 9 regions was reduced by at least 5% between 1981-2010 and 2015-2019. {12.3.1, 12.3.2, 12.3.6, Table 12.4} 10 11 Climate change affects the epidemiology of climate-sensitive infectious diseases in the region (high 12 confidence). Examples are the effects of warming temperatures on increasing the suitability of transmission 13 of vector-borne diseases, including endemic and emerging arboviral diseases such as dengue fever, 14 chikungunya, and Zika (medium confidence). The reproduction potential for the transmission of dengue 15 increased between 17% and 80% for the period 1950-54 to 2016-2021 depending on the subregion as a result 16 of changes in temperature and precipitation (high confidence). {12.3.1, 12.3.2, 12.3.3, 12.3.5, 12.3.6, Table 17 12.1} 18 19 The Andes, northeast Brazil and the northern countries in Central America are among the more 20 sensitive regions to climatic-related migrations and displacements, a phenomenon that has increased 21 since AR5 (high confidence). Climatic drivers interact with social, political, geopolitical and economical 22 drivers; the most common climatic drivers for migration and displacements are droughts, tropical storms and 23 hurricanes, heavy rains and floods (high confidence). {12.3.1.4, 12.3.2.4, 12.3.3.4, 12.3.5.4, 12.5.8.4} 24 25 The impacts of climate change are not of equal scope for men and women (high confidence). Women, 26 particularly the poorest, are more vulnerable and are impacted in greater proportion. Often they have less 27 capacity to adapt, further widening structural gender gaps (high confidence). {12.3.7.3, 12.5.2.4, 12.5.2.5, 28 12.5.7.3, 12.5.8.1, 12.5.8.3, 12.5.8.4} 29 30 Current adaptation responses: 31 32 Ecosystem-based adaptation is the most common adaptation strategy for terrestrial and freshwater 33 ecosystems (high confidence). There is a focus on the protection of native terrestrial vegetation through 34 implementation of protected areas and payment for ecosystem services, especially those related to water 35 provision. The adaptation measures in place, however, are insufficient to safeguard terrestrial and freshwater 36 ecosystems in the CSA from negative impacts of climate change (high confidence). {12.5.1, 12.5.3, 12.6} 37 38 Adaptation initiatives in ocean and coastal ecosystems mainly focus on conservation, protection and 39 restoration) (high confidence). The main adaptation measures are ocean zoning, the prohibition of 40 productive activities (e.g., fisheries, aquaculture, mining, tourism) on marine ecosystems, the improvement 41 of research and education programs, and the creation of specific national policies (high confidence). {12.5.2} 42 43 Adaptive water management has mainly centred on enhancing quantity and quality of water supply, 44 including large infrastructure projects, which, however, are often contested and can exacerbate water 45 related conflicts (high confidence). Inclusive water regimes that overcome social inequalities and 46 approaches including nature-based solutions, such as wetland restoration and water storage and infiltration 47 infrastructure, with synergies for ecosystem conservation and disaster risk reduction, have been found to be 48 more successful for adaptation and sustainable development (high confidence). {12.5.3, 12.6.1, 12.6.3} 49 50 Adaptation strategies for agricultural production are increasing in the region as a response to current 51 and projected changes in climate (high confidence). The main observed adaptation strategies in agriculture 52 and forestry are soil and water management conservation, crop diversification, climate-smart agriculture, 53 early warning systems, upward shifting for plantations to avoid warming habitat and pests and improved 54 management of pastures and livestock. Adaptation requires governance improvements and new strategies to 55 address changing climate; nevertheless, barriers limiting adaptive capacity persist such as lack of educational 56 programs for farmers, adequate knowledge of site-specific adaptation and institutional and financial 57 constraints (high confidence). {12.5.4} Do Not Cite, Quote or Distribute 12-4 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Urban adaptation in the region includes solutions on regulation, planning, urban waters management 3 and housing (high confidence). Regulation, planning and control systems are central tools on reducing risk 4 associated with the security of the buildings, their location, and the proper supply of basic urban services and 5 transport (high confidence). The adoption of nature-based solutions (e.g., urban agriculture and rivers 6 restoration) and hybrid (grey-green) infrastructure are still incipient with weak connections to poverty and 7 inequality reduction strategies (medium confidence). Focusing on risk reduction encompasses upgrading 8 informal and precarious settlements, built-environments, and improving housing conditions, which offer an 9 important but still limited contribution to urban adaptation (high confidence). {12.5.5, 12.5.7, 12.6.1} 10 11 Adaptation initiatives for the health sector are mainly focused on the development of climate services 12 such as integrated climate-health surveillance and observatories, forecasting climate-related disasters 13 and vulnerability maps (high confidence). Climate services for the health sector are largely focused on 14 epidemic forecast tools and associated early warning systems for vector-borne diseases and heat and cold 15 waves. Political, institutional and financial barriers reduce the feasibility of implementing these tools (high 16 confidence). {12.5.6, Table 12.9, Table 12.11} 17 18 Indigenous knowledge and local knowledge are crucial for the adaptation and resilience of social- 19 ecological systems (high confidence). Indigenous knowledge and local knowledge can contribute to 20 reducing the vulnerability of local communities to climate change (medium confidence). {12.5.1, 12.5.8, 21 12.6.2} 22 23 What are the projected impacts and key risks? 24 25 Climate change is projected to convert existing risks in the region into severe key risks (medium 26 confidence). Key risks are assessed as follows: 1. Risk of food insecurity due to droughts; 2. Risk to people 27 and infrastructure due to floods and landslides; 3. Risk of water insecurity due to declining snow cover, 28 shrinking glaciers and rainfall variability; 4. Risk of increasing epidemics particularly of vector-borne 29 diseases; 5. Cascading risks surpassing public service systems; 6. Risk of large-scale changes and biome 30 shifts in the Amazon; 7. Risks to coral reef ecosystems; and 8. Risks to coastal socio-ecological systems due 31 to sea level rise, storm surges and coastal erosion. {12.3, 12.4, Figure 12.9, Figure 12.11, Table 12.6, Table 32 SM12.5} 33 34 Impacts on rural livelihoods and food security, particularly for small and medium-sized farmers and 35 Indigenous Peoples in the mountains, are projected to worsen, including the overall reduction of 36 agricultural production, suitable farming area and water availability (high confidence). Projected yield 37 reductions by 2050 under A2 scenario are: bean 19%, maize 4­21%, rice 23% in Central America with 38 seasonal droughts projected to lengthen, intensify and increase in frequency. Small fisheries and farming of 39 seafood will be negatively affected as ENSO events become more frequent and intense and ocean warming 40 and acidification continues (medium confidence). {12.2, 12.3, 12.4, Figure 12.9, Figure 12.11, Table 12.4} 41 42 Extreme precipitation events, which result in floods, landslides and droughts, are projected to 43 intensify in magnitude and frequency due to climate change (medium confidence). Floods and landslides 44 pose a risk to life and infrastructure; a 1.5ºC increase would result in an increase of 100­200% in the 45 population affected by floods in Colombia, Brazil and Argentina, 300% in Ecuador and 400% in Peru 46 (medium confidence). {12.3, Figure 12.7, Figure 12.9, Table SM12.5} 47 48 Increasing water scarcity and competition over water are projected (high confidence). Disruption in 49 water flows will significantly degrade ecosystems such as high-elevation wetlands and affect farming 50 communities, public health and energy production (high confidence). {12.3, Figure 12.3, Figure 12.9, Figure 51 12.11} 52 53 In the next decades, endemic and emerging climate-sensitive infectious diseases are projected to 54 increase (medium confidence). This can happen through expanded distribution of vectors, especially viral 55 infectious diseases from zoonotic origin in transition areas between urban and suburban, or rural settings, 56 and upslope in the mountains (medium confidence). {12.3.2, 12.3.5, 12.3.7, Figure 12.5, Figure 12.9, Figure 57 12.11, Table 12.6, Table SM12.5} Do Not Cite, Quote or Distribute 12-5 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 The positive feedback between climate change and land use change, particularly deforestation, is 3 projected to increase the threat to the Amazon forest, resulting in the increase of fire occurrence, 4 forest degradation (high confidence) and long-term loss of forest structure (medium confidence). The 5 combined effect of both impacts will lead to a long-term decrease in carbon stocks in forest biomass, 6 compromising Amazonia´s role as a carbon sink, largely conditional on the forest's responses to elevated 7 atmospheric CO2 (medium confidence). The southern portion of the Amazon has become a net carbon source 8 to the atmosphere in the past decade (high confidence). {12.3.3, 12.3.4, Figure 12.9, Figure 12.11, Table 9 12.6, Table SM12.5} 10 11 Up to 85% of natural systems (plant and animal species, habitats and communities) evaluated in the 12 literature for biodiversity-rich spots in the region are projected to be negatively impacted by climate 13 change (medium confidence). Available studies focus mainly on vertebrates and plants of the Atlantic 14 Forest and Cerrado in Brazil and in Central America, with a large knowledge gap on freshwater ecosystems 15 {12.3, 12.5.1, CCP1} 16 17 Ocean and coastal ecosystems in the region will continue to be highly impacted by climate change 18 (high confidence). Coral reefs are projected to lose their habitat, change their distribution range and suffer 19 more bleaching events driven by ocean warming. In the RCP4.5 and RCP8.5 scenarios by 2050, virtually 20 every coral reef will experience at least one severe bleaching event per year (high confidence). Under all 21 RCP scenarios of climate change, there will be changes in the geographical distribution of marine species 22 and ocean and coastal ecosystems such as mangroves, estuaries, rocky shores, as well as those species 23 subjected to fisheries (medium confidence). {Figure 12.9, Table SM12.3, Table SM12.4} 24 25 Contribution of adaptation to solutions and barriers to adaptation 26 27 Policies and actions at multiple scales and the participation of actors from all social groups, including 28 the most exposed and vulnerable populations, are critical elements for effective adaptation (high 29 confidence). Engaging social movements and local actors in policy-making and planning for adaptation 30 generates positive synergies and better results. Adaptation policies and programs that consider age, 31 socioeconomic status, race, and ethnicity are more efficient, as these factors determine vulnerability and 32 potential benefits of adaptation. Socio-economic and political factors that provide some level of safety and 33 continuity of policies and actions are critical enablers of adaptation (high confidence). {12.5.1, 12.5.2, 34 12.5.7, 12.5.8, 12.6.4} 35 36 The knowledge and awareness of climate change as a threat has been increasing since AR5 due to the 37 increasing frequency and magnitude of extreme weather events in the region, information available 38 and climate justice activism (high confidence). Conflicts in which direct biophysical impacts of climate 39 change play a major role can unleash protests and strengthen social movements (medium confidence). 40 {12.5.8, 12.6.4} 41 42 Research approaches that integrate Indigenous knowledge and local knowledge systems, with natural 43 and social sciences, have increased since AR5 (high confidence), and are helping to improve decision- 44 making processes in the region, reduce maladaptation, and foster transformational adaptation through the 45 integration with ecosystem-based adaptation and community-based adaptation (high confidence). {12.5.1, 46 12.5.8, 12.6.2} 47 48 The most reported obstacle for adaptation in terrestrial, freshwater, ocean and coastal ecosystems is 49 financing (high confidence). There is also a significant gap in identifying limits to adaptation and weak 50 institutional capacity for implementation. This hinders the development of comprehensive adaptation 51 programs, even under adequate funding. {12.5.1, 12.5.2} 52 53 Climate Smart Agriculture technologies strengthening synergies among productivity and mitigation is 54 growing as an important adaptation strategy in the region (high confidence). Pertinent information for 55 farmers provided by Climate Information Services are helping them to understand the role of climate vs. 56 other drivers in perceived productivity changes. Index insurance builds resilience and contributes to Do Not Cite, Quote or Distribute 12-6 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 adaptation both by protecting farmers' assets in the face of major climate shocks, by promoting access to 2 credit, and by the adoption of improved farm technologies and practices. {12.5.4} 3 4 Institutional instability, fragmented services and poor water management, inadequate governance 5 structures, insufficient data and analysis of adaptation experience are barriers to address the water 6 challenges in the region (high confidence). {12.5.3} 7 8 Inequality, poverty and informality shaping cities in the region increase vulnerability to climate 9 change while policies, plans or interventions addressing these social challenges with inclusive 10 approaches are opportunities for adaptation (high confidence). Initiatives to improve informal and 11 precarious settlement, guaranteeing access to land and decent housing, are aligned with comprehensive 12 adaptation policies that include development and reduction of poverty, inequality and disaster risk (medium 13 confidence). {12.5.5, 12.5.7} 14 15 Adaptation policies often address climate impact drivers, but seldom include the social and economic 16 underpinnings of vulnerability. This narrow scope limits adaptation results and compromises their 17 continuity in the region (high confidence). In a context of unaddressed underdevelopment, adaptation 18 policies tackling poverty and inequality are marginal, underfunded, and not clearly included at national, 19 regional or urban levels. Dialogue and agreement including multiple actors are mechanisms to acknowledge 20 trade-offs and promote dynamic, site-specific adaptation options (medium confidence). {12.5.7} 21 22 Do Not Cite, Quote or Distribute 12-7 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 12.1 Introduction 2 3 12.1.1 The Central and South America Region 4 5 Central and South America (CSA) is a highly diverse region, both culturally and biologically. It harbours one 6 of the highest biodiversity on the planet (Hoorn et al., 2010; Zador et al., 2015; IPBES, 2018a) (Cross- 7 Chapter Paper 1: Biodiversity Hotspots) and a wealth of cultural diversity resulting from more than 800 8 Indigenous Peoples who share the territory with European and African descendants and more recent Asian 9 migrants (CEPAL, 2014). Moreover, it is one of the most urbanized regions in the world, with some of the 10 most populated metropolitan areas (UNDESA, 2019). Several countries in the region have experienced 11 sustained economic growth in the last decades, making important advances in reducing poverty in the area. 12 Yet, it is a region of substantial social inequality including the highest inequality in land tenure, where there 13 still remains a large percentage of the population below the poverty line, unequally distributed between rural 14 and urban areas and along aspects like gender and race; these groups are highly vulnerable to climate change 15 and natural extreme events that frequently affect the region (high confidence) (ECLAC, 2019b; Busso and 16 Messina, 2020; Poveda et al., 2020). 17 18 Land use changes in the region, particularly deforestation, are large, mostly due to agricultural production for 19 export purposes, one of the main sources of income for the area (Salazar et al., 2016) (Figure 12.2c). 20 Additional pressure on the land comes from illegal activities, pollution and induced fires. These changes 21 exacerbate the impacts of climate change and make the region play a key role in the future of the world 22 economy and food production (IPBES, 2018a). The region boasts the largest tropical forest on the planet and 23 other important biomes of high biodiversity on mountains, lowlands and coastal areas. It can potentially 24 continue its agricultural expansion and development at the expense of substantially reducing the areas of 25 natural biomes. Indigenous Peoples and smallholder families are lacking adequate climate policies combined 26 with institutions to protect their property rights; this could result in a more sustainable process of agricultural 27 expansion, without substantially increasing greenhouse gas emissions and the vulnerability of those 28 populations (high confidence) (Sá et al., 2017). 29 30 Central and South America (CSA) is divided into eight climatic sub-regions by WGI (Figure 12.1). Though 31 the southern part of Mexico is included in the climatic sub-region SCA for WGI, Mexico is assessed in 32 Chapter 14 (North America). In this chapter, we refer to this sub-region as Central America (CA) as it 33 excludes southern Mexico. The climate change literature for the region occasionally includes Mexico and in 34 those cases, our assessment makes reference to Latin America but when only southern Mexico is included, 35 the term Mesoamerica is used. Figure 12.2 and Table SM12.1 summarize relevant characteristics of the sub- 36 regions included in this chapter. 37 38 Do Not Cite, Quote or Distribute 12-8 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Figure 12.1: Sub-regions included in the Central and South America region. Note that the WGI climatic sub-region 3 South Central America SCA corresponds to Central America CA in this chapter, as southern Mexico is included in 4 Chapter 14. Small islands in the region are approached in Chapter 15 in more detail. 5 Do Not Cite, Quote or Distribute 12-9 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Figure 12.2: Characterization of the region. Population data from ISIMIP (2021) after Klein Goldewijk et al. (2017). 3 Biodiversity expressed as marine and terrestrial species richness adapted from Gagné et al. (2020). Land cover data 4 from ESA (2018). Human Development Index and its components from UNDP (2020). HDI and components for 5 French Guiana from Global Data Lab (2020). 6 7 8 12.1.2 Approach and Storyline for the Chapter 9 10 The chapter is divided in two main sections. The first section follows an integrative approach in which 11 hazards, exposure, vulnerability, impacts and risks are discussed following the eight climatically 12 homogeneous sub-regions described in WGI AR6 (see Figure 12.1). The second section assesses the 13 implemented and proposed adaptation practices by sector; in doing so, it connects to the WGII AR6 cross- 14 chapters themes. The storyline is then a description of the hazards, exposure, vulnerability and impacts 15 providing as much detail as available in the literature at the sub-regional level, followed by the identification 16 of risks as a result of the interaction of those aspects. This integrated sub-regional approach ensures a 17 balance in the text, particularly for countries that are usually underrepresented in the literature but that show 18 a high level of vulnerability and impacts, such as those observed in CA. The sectoral assessment of 19 adaptation that follows is useful for policy makers and implementers, usually focused and organized by 20 sectors, governments' ministries or secretaries that can easily locate the relevant adaptation information for 21 their particular sector. To ensure coherence in the chapter, a summary of the assessed adaptation options by Do Not Cite, Quote or Distribute 12-10 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 key risks is presented, followed by a feasibility assessment for some relevant adaptation options. The chapter 2 closes with case studies and a discussion of the knowledge gaps evidenced in the process of the assessment. 3 4 5 12.2 Summary of the Fifth Assessment Report and Recent IPCC Special Reports 6 7 Central and South America shows increasing trends of climatic change and variability and extreme events 8 severely impacting the region, exacerbating problems of rampant and persistent poverty, precarious health 9 systems and water and sanitation services, malnutrition and pollution. Inadequate governance and lack of 10 participation escalates the vulnerability and risk to climate variability and change in the region (high 11 confidence) (WGII AR5 Chapter 27) (Magrin et al., 2014). 12 13 Increasing trends in precipitation had been observed in Southeast South America (SES in Figure 12.1) in 14 contrast with decreasing trends in CA and central-southern Chile (high confidence) (WGII AR5 Chapter 27) 15 (Magrin et al., 2014). Frequency and intensity of droughts have increased in many parts of SA (IPCC, 16 2019c). Warming has been detected throughout CSA except for a cooling trend reported for the ocean off the 17 Chilean coast. 18 19 Climate projections indicate increases in temperature for the entire region by 2100 for RCP4.5 and RCP8.5, 20 but rainfall changes will vary geographically, with a notable reduction of ­22% in Northeast Brazil and an 21 increase of +25% in SES. Significant dependency on rainfed agriculture (>30% in Guatemala, Honduras, and 22 Nicaragua) indicates high sensitivity to climatic variability and change, and challenge food security (high 23 confidence) (SRCCL Chapter 5, Mbow et al., 2019). Undernutrition has worsened since 2014 in CSA 24 (SRCCL Chapter 5, Mbow et al., 2019). Evidence of climate change impacts on food security is emerging 25 from Indigenous knowledge and local knowledge studies in SA. Municipalities in CA with high proportion 26 of subsistence crops tend to have less resources for adaptation and more vulnerable to climate change 27 (SRCCL Chapter 5, Mbow et al., 2019). Rising temperature and decreased rainfall could reduce agricultural 28 productivity by 2030, threatening food security of the poorest populations (WGII AR5 Chapter 27, Magrin 29 et al., 2014). Though reduced suitability and yield for beans, coffee, maize, plantain, and rice is expected in 30 CA (SRCCL Chapter 5, Mbow et al., 2019), limiting the warming to 1.5ºC, compared with 2ºC, is projected 31 to result in smaller net reductions in yields of maize, rice, wheat and other cereal crops for CSA (high 32 confidence) (SR15 Chapter 3, Hoegh-Guldberg et al., 2018). The heat stress is expected to reduce the 33 suitability of Arabica coffee in Mesoamerica but it can improve in high latitude areas in SA (SRCCL 34 Chapter 4, Olsson et al., 2019). There is limited evidence that these declines in crop yields may result in 35 significant population displacement from the tropics to the subtropics (SR15 Chapter 3, Hoegh-Guldberg et 36 al., 2018). 37 38 There is a high confidence that heat waves will increase in frequency, intensity and duration, becoming, 39 under high emission scenarios, extremely long, over 60 days in duration in SA; the risk of wildfires will also 40 increase significantly in SA (SRCCL Chapter 2, Jia et al., 2019). These processes are and will lead to 41 increased desertification that cost between 8 and 14% of gross agricultural product in many CSA countries 42 (SRCCL Chapter 3, Mirzabaev et al., 2019). Distinguishing climate induced changes from land use changes 43 is challenging, but 5­6% of biomes in SA are expected to change by 2100 due to climate change (SRCCL 44 Chapter 4, Olsson et al., 2019). 45 46 Changes in weather and climatic patterns are negatively affecting human health in CSA, in part through the 47 emergence of diseases in previously non-endemic areas (WGII AR5 Chapter 27, Magrin et al., 2014). 48 Projections of potential impacts of climate change on malaria confirm that weather and climate are among 49 the drivers of geographic range, intensity of transmission, and seasonality; the changes of risk become more 50 complex with additional warming (very high confidence) (SR15 Chapter 3, Hoegh-Guldberg et al., 2018). 51 There is high confidence that constraining the warming to 1.5°C would reduce risks for unique and 52 threatened ecosystems safeguarding the services they provide for livelihoods and sustainable development 53 (food, water) in CA and Amazon (SR15 Chapter 5, Roy et al., 2018). 54 55 Observed changes in streamflow and water availability affect vulnerable regions (WGII AR5 Chapter 27, 56 Magrin et al., 2014). Glacier mass changes in the Andes over the past decades are among the most negative 57 ones worldwide (SROCC Chapter 2, Hock et al., 2019). This reduction has modified the frequency, Do Not Cite, Quote or Distribute 12-11 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 magnitude and location of related natural hazards, while the exposure of people and infrastructure has 2 increased because in relation with growing population, tourism and economic development (high confidence) 3 (SROCC Chapter 2, Hock et al., 2019). 4 5 Negative impacts of climate change in the region are exacerbated by deforestation and land degradation 6 attributed mainly to expansion and intensification of agriculture and cattle ranching, usually under insecure- 7 tenure land. This conversion of natural ecosystems is the main cause of biodiversity and ecosystem loss and 8 is an important source of greenhouse gas (GHG) emissions (high confidence) (WGII AR5 Chapter 27, 9 Magrin et al., 2014). 10 11 The combination of continued anthropogenic disturbance, particularly deforestation, with global warming 12 may result in dieback of forest in the region (medium confidence) (SR15 Chapter 3, Hoegh-Guldberg et al., 13 2018). Loses as high as 40% of biomass are projected in CA with a warming of 3°C­4°C and the Amazon 14 may experience a significant dieback at similar warming levels (SR15 Chapter 3, Hoegh-Guldberg et al., 15 2018). Advances in second-generation bioethanol from sugarcane and other feedstock will be important for 16 mitigation. However, agricultural expansion results in large conversions in tropical dry woodlands and 17 savannas in SA (Brazilian Cerrado, Caatinga and Chaco) (high confidence) (SRCCL Chapter 1, Arneth et al., 18 2019). The expansion of soybean plantations in the Amazonian state of Mato Grosso in Brazil reached 19 16.8% yr-1 from 2000 to 2005; and oil palm, a significant biofuel crop, is also linked to recent deforestation 20 in tropical CA (Costa Rica and Honduras) and SA (Colombia and Ecuador), although lower in magnitude 21 compared to deforestation from soybean and cattle ranching (WGII AR5 Chapter 27, Magrin et al., 2014). 22 23 Ocean and coastal ecosystems in the region already show important changes due to climate change and 24 global warming (SROCC Chapter 5, Bindoff et al., 2019). 25 26 Adaptation to future climate changes starts by reducing the vulnerability to present climate considering the 27 deficient welfare of people in the region. Generalizing to the region cases of synergies among development, 28 adaptation and mitigation planning requires a governance model where development needs, vulnerability 29 reduction, and adaptation strategies are intertwined (WGII AR5 Chapter 27, Magrin et al., 2014). 30 31 32 12.3 Hazards, Exposure, Vulnerabilities and Impacts 33 34 12.3.1 Central America (CA) Sub-region 35 36 12.3.1.1 Hazards 37 38 Since the mid-20th century, extreme warm temperatures have increased and extreme cold temperatures have 39 decreased in the region (medium confidence). The magnitude and frequency of extreme precipitation events 40 have increased, but droughts have mixed signals (low confidence) (WGI AR6 Table 11.13, Table 11.14, 41 Table 11.15, Seneviratne et al., 2021). There are spatially variable trends detected for the mid-summer 42 drought (MSD) timing, the amount of rainy season precipitation, the number of consecutive and total dry 43 days, and extreme wet events at the local scale since the 1980s. At the regional scale, a positive trend in the 44 duration, but not the magnitude of the MSD was found (Anderson et al., 2019). 45 46 Significant increases in tropical cyclone (TC) intensification rates in the Atlantic basin, highly unusual 47 compared to model-based estimates of internal climate variations has been observed (Bhatia et al., 2019). TC 48 contributed approximately 10% of the annual precipitation (Khouakhi et al., 2017). During the TC season 49 more TC-driven events of extreme sea level exceed a 10-year return period (Muis et al., 2019). 50 51 Massive heat wave events and increase in the frequency of warm extremes are projected at the end of the 52 21st century (high confidence). When comparing 2.0 with 1.5 degrees of warming, the longest annual warm 53 wave is projected to increase more than 60 days (Taylor et al., 2018). 54 55 General decrease in the magnitude of heavy precipitation extremes (Chou et al., 2014; Giorgi et al., 2014) (in 56 1.5ºC projection) but increase in the frequency of extreme precipitation (R50mm) (Imbach et al., 2018) are 57 projected for both 2ºC and 4ºC GWL. Strong declines in mean daily rainfall are projected for July in Belize Do Not Cite, Quote or Distribute 12-12 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 (Stennett-Brown et al., 2017; WGI AR6 Table 11.14, Seneviratne et al., 2021) and decreased rainfall through 2 the year for all capital cities except Panama City (medium confidence: limited evidence, high agreement) 3 (Pinzón et al., 2017). 4 5 The main climate impact drivers like extreme heat, drought, relative sea level rise, coastal flooding, erosion, 6 marine heatwaves, ocean aridity, (high confidence) and aridity, drought and wildfires will increase by mid- 7 century (medium confidence) (Figure 12.6, WGI AR6 Table 12.6, Ranasinghe et al., 2021). 8 9 The rainy season in CA will likely experience more pronounced MSD by the end of this century, with a 10 signal for reduced minimum precipitation by the mid-century for the JJA and SON quarters, and a broader 11 second peak is projected consistent with the future south displacement of the ITCZ (high confidence) 12 (Fuentes-Franco et al., 2015; Hidalgo et al., 2017; Maurer et al., 2017; Imbach et al., 2018; Naumann et al., 13 2018; Ribalaygua et al., 2018; Corrales-Suastegui et al., 2020). 14 15 Climate projections indicate a decrease in frequency of tropical cyclones in CA accompanied with an 16 increased frequency of intense cyclones (WGI AR6 Section 12.4.4.3, Ranasinghe et al., 2021). 17 18 12.3.1.2 Exposure 19 20 Of the 47 million Central Americans in 2015, 40% lived in rural areas with Belize being the least urbanized 21 (54% rural) and Costa Rica the most (21% rural) (CELADE, 2019); 10.5 million lived in the Dry Corridor 22 region, an area recently exposed to severe droughts that have resulted in 3.5 million people in need of 23 humanitarian assistance (FAO, 2016a). Except in Belize and Panama, the majority of the countries' 24 population --ranging from 56% in Honduras to 95% in El Salvador-- is exposed to 2 or more risks derived 25 from natural extreme events, affecting between 57% to 96% of the GDP of the countries (UNISDR and 26 CEPREDENAC, 2014). Central America is one of the regions most exposed to climatic phenomena; with 27 long coastlines and lowland areas, the region is repeatedly affected by drought, intense rains, cyclones and 28 ENSO events (high confidence) (ECLAC et al., 2015). 29 30 Large urban centres are located on mountains or away from the shore, with the notable exceptions of Panama 31 City, Belmopan and Managua, capital cities housing around 3 million people. Urban development in the 32 capital cities and suburbs has almost tripled in the last forty years reaching population densities as high as 33 11,000 inhabitants per km2 in Guatemala City and Tegucigalpa, with the spread of poor neighbourhoods in 34 steep ravines and other marginal high risk areas (Programa Estado de la Nación - Estado de la Región, 2016). 35 36 12.3.1.3 Vulnerability 37 38 Climate change is exacerbating socioeconomic vulnerability in CA, a region with high levels of 39 socioeconomic, ethnic and gender inequality, high rates of child and maternal mortality and morbidity, high 40 levels of malnutrition and inadequate access to food and drinking water (ECLAC et al., 2015). Disasters 41 from adverse natural events exacerbate CA's economic vulnerability, accounting for substantial human and 42 economic losses (UNISDR and CEPREDENAC, 2014). Vulnerability in most sectors is considered high or 43 very high (high confidence) (Figure 12.7). 44 45 Approximately 40% of the CA population are living in poverty. Guatemala (62%), Honduras (60%), 46 Nicaragua (46%) and Belize (42%, 2009) had the highest poverty rates in CSA in 2018 (ECLAC, 2019b; 47 BCIE, 2020). Rural poverty rates are higher, 82% in Honduras and 77% in Guatemala in 2014, and so is 48 poverty among Indigenous Peoples, up to 79% in Guatemala. Rural poor are the most sensitive to climate 49 extremes as their main economic activity is based on agriculture in vulnerable terrains (NU CEPAL, 2018). 50 In 2014, all CA countries, except for El Salvador (excluding Belize), had higher GINI coefficients (more 51 inequality) than the average for Latin America (0.473), which in itself is the most unequal region in the 52 world (ECLAC, 2019b); in 2018 the situation remained similar with El Salvador showing the lowest GINI 53 coefficient (40) and the rest of the countries showing values higher than the Latin-American average (BCIE, 54 2020). 55 56 12.3.1.4 Impacts 57 Do Not Cite, Quote or Distribute 12-13 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 The countries in the region are consistently ranked with the highest risk in the world of being impacted by 2 extreme events (high confidence). Economic cost of climate change impacts in 2010 was estimated from 3 2.9% of GDP for Guatemala to 7.7% for Belize (ECLAC et al., 2015). For the period 1992­2011, Honduras, 4 Nicaragua and Guatemala were among the 10 most impacted countries in the world by extreme weather 5 events (UNISDR and CEPREDENAC, 2014). The number of these events has increased 3% annually in the 6 last 30 years (Bárcena et al., 2020a). 7 8 Human and economic losses, changing water availability and increasing food insecurity are the most studied 9 impacts of climate change in CA (Figure 12.9; Harvey et al., 2018; Hoegh-Guldberg et al., 2019). Hydro- 10 meteorological events, such as storm surges and tropical cyclones, are the most frequent extreme events and 11 have the highest impact (high confidence) (Reyer et al., 2017). From 2005 to 2014, the cumulative impacts 12 were over 3410 people dead, hundreds of thousands displaced, and damages estimated around USD 5.8 13 billion (Ishizawa and Miranda, 2016). One standard deviation in the intensity of a hurricane windstorm leads 14 to a decrease in both the growth of total GDP per capita (0.9% to 1.6%) and total income and labour income 15 by 3%, whereas it increases moderate and extreme poverty by 1.5% in CA (Ishizawa and Miranda, 2016). 16 17 Food insecurity is a serious impact of climate change in a region where 10% of the GDP depends on 18 agriculture, livestock and fisheries (very high confidence) (ECLAC et al., 2015; CEPAL et al., 2018; Harvey 19 et al., 2018; BCIE, 2020). Crop losses largely result from highly variable rainfall and seasonal droughts 20 which have increased significantly in the last decades (Table 12.3; CEPAL and CAC-SICA, 2020), 21 particularly the observed changes in the MSD that reduces rainfall at the onset of the rainy season (May- 22 June) (Anderson et al., 2019). Small and subsistence farmers receive the highest impact as they practice 23 rainfed agriculture (Imbach et al., 2017), and poor neighbourhoods, which face socioeconomic and physical 24 barriers for adapting to climate change (Kongsager, 2017). In 2015, precipitation diminished between 50% to 25 70% of its historic average causing the loss of up to 80% of beans and 60% of maize, leaving 2.5 million 26 people food insecure, 1.6 million of which were in the Dry Corridor of CA (ECLAC et al., 2015; FAO, 27 2016a). In 2019, the region entered its fifth consecutive drought year with 1.4 million people in need of food 28 aid. Seasonal-scale droughts are projected to lengthen by 12­30%, intensify by 17­42% and increase in 29 frequency by 21­42% in RCP4.5 and RCP8.5 scenarios by the end of the century (Depsky and Pons, 2021). 30 31 Studies have shown that the incidence of some vector-borne and zoonotic diseases in CA is correlated to 32 climatic variables, particularly temperature and rainfall (high confidence) (Figure 12.4; Table 12.1). In 33 Honduras, rainfall and relative humidity were positively correlated with the occurrence of hemorrhagic 34 dengue cases (Zambrano et al., 2012). In Costa Rica, temperature and rainfall was correlated to cattle rabies 35 outbreaks and mortality during 1985­2016 (Hutter et al., 2018); Incidence of leishmaniasis showed cycles of 36 three years related to temperature changes (Chaves and Pascual, 2006); and snakebites were more likely to 37 occur at high temperatures and was significantly reduced after the rainy season for the period 2005­2013 38 (Chaves et al., 2015). In Panama, rainfall was associated with the increased number of malaria cases among 39 the Gunas, an Indigenous People with high vulnerability living in poverty conditions on small islands 40 affected by sea-level rise (Hurtado et al., 2018). These correlations point to a possible change in disease 41 incidence with climate change; evidence of that change is yet to be reported in the literature as longitudinal 42 studies are lacking in the region. 43 44 Heat stress is another health concern in this already warm and humid part of the world (high confidence) 45 (Table 12.2); it is an increasing occupational health hazard with potential impacts on kidney disease 46 (Sheffield et al., 2013; Dally et al., 2018; Johnson et al., 2019). Sea-level rise exacerbating wave-driven 47 flooding is expected to impact infrastructure and freshwater availability in small islands and atolls off the 48 coast of Belize (Storlazzi et al., 2018). Observed and expected impacts in the coastal and ocean ecosystems 49 of the sub-region are described in Figure 12.9. 50 51 Decreasing water availability is another impact of climate change (high confidence). Under a climate change 52 scenario of 3.5°C warming and a 30% reduction of rainfall, a reduction in production and export of crops and 53 livestock is projected affecting the wages and decreasing the GDP of Guatemala by 1.2%, thereby increasing 54 food insecurity (Vargas et al., 2018b). By 2100, water availability per capita is projected to decrease 82% 55 and 90% on average for the region under B2 (low emissions) and A2 (high emissions) scenarios respectively 56 (CEPAL, 2010) (Figure 12.3). 57 Do Not Cite, Quote or Distribute 12-14 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Figure 12.3: Reduction of water availability per capita projected to 2100 without climate change (baseline scenario) 3 and with two climate change scenarios (CEPAL, 2010). 4 5 6 Impacts on rural livelihoods, particularly for small and medium-sized farmers and Indigenous Peoples on the 7 mountains, include the overall reduction of the production, yield (Table 12.4), suitable farming area, and 8 water availability (high confidence) (Walshe and Argumedo, 2016; Bouroncle et al., 2017; Hannah et al., 9 2017; Imbach et al., 2017; Harvey et al., 2018; Batzín, 2019; Donatti et al., 2019). Bean production in El 10 Salvador, Nicaragua, Honduras, and Guatemala, is projected to decrease, using the Decision Support for 11 Agro-Technology Transfer (DSSAT) under A2 scenario, by 19% for 2050, whereas maize production, 12 depending on water retention capacity of soils, will drop between 4% and 21% by 2050 (CEPAL et al., 13 2018). In Guatemala, the yield of rainfed maize is expected to decrease by 16% by 2050 under RCP8.5 using 14 the Global Gridded Crop Model Intercomparison GGCMI; yields for rainfed sugarcane are expected to drop 15 by 44% and irrigated sugarcane by 36% under the same modelling conditions (Castellanos et al., 2018). Rice 16 production is expected to decrease by 23% under scenario A2 by 2050 (CEPAL and CAC/SICA, 2013). 17 18 The extent and quality of suitable areas for basic grains are expected to contract (high confidence). The 19 suitable area for maize will experience a 35% reduction of cultivated area expected by 2100 under A2 20 scenario. The area suitable for beans is expected to reduce by 2050. Projections show that suitable areas with 21 excellent aptitude under current conditions will decrease by 14%, mainly in Panama (41%) Costa Rica (21%) 22 and El Salvador (20%). Species Distribution Model, using the IPSL GCM, projects that the suitable zones 23 for cacao and coffee will shrink between 25% to 75% under RCP6.0 (Fernandez-Manjarrés, 2018; Fernández 24 Kolb et al., 2019). Warmer and dryer lower areas will become unsuitable for coffee and will drive its 25 production to higher land (Läderach et al., 2013; Bunn et al., 2015). Under A2 climate change scenario, areas 26 with excellent aptitude for Arabica coffee will decrease by 12% in Central America; coffee yield will 27 decrease in suitable zones whereby the extent of high yield (> 0.8 T ha-1) zones is project to shrink from 34% 28 to 12% whereas low yield (< 0.3 T ha-1) zone will expand from 14% to 36% by 2100 under A2 scenario 29 (CEPAL and CAC/SICA, 2014). 30 31 The Mesoamerica, biodiversity-rich spot spanning through CA and southern Mexico is a global priority for 32 terrestrial biodiversity conservation, and it is projected to be negatively impacted by climate change, 33 especially through the contraction of distribution of native species at the area becomes increasingly dryer 34 (high confidence) (Cross-Chapter Paper 1.2.2; Feeley et al., 2013; Manes et al., 2021) . A significant 35 reduction in net primary productivity in tropical forests is expected under both RCP4.5 and RCP8.5 as a 36 result of temperature increase, precipitation reduction, and droughts (Lyra et al., 2017; Castro et al., 2018; 37 Stan et al., 2020). Models of aridity index show that the dry, sub humid vegetation of the dry corridor will 38 expand to neighbouring areas and replace the humid forests in the Pacific lowlands and the northern parts of 39 Guatemala by 2050 under RCP4.5 and RCP8.5 scenarios (Pons et al., 2018; CEPAL and CAC-SICA, 2020). Do Not Cite, Quote or Distribute 12-15 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 3°C warming would shrink the tropical rainforest and replace it with savannah grassland. Wetlands are also 2 expected to be highly affected by climate change in the region (Hoegh-Guldberg et al., 2019). 3 4 12.3.2 Northwest South America (NWS) Sub-region 5 6 12.3.2.1 Hazards 7 8 Significant increases in the intensity and frequency of hot extremes and significant decreases in the intensity 9 and frequency of cold extremes (Dereczynski et al., 2020; Dunn et al., 2020) was likely2 observed (Figure 10 12.6; WGI AR6 Table 11.13, Seneviratne et al., 2021). 11 12 Insufficient data coverage and trends in available data are generally not significant for heavy precipitation 13 (low confidence) (Dereczynski et al., 2020; Dunn et al., 2020; Sun et al., 2021) (Figure 12.6; WGI AR6 14 Table 11.14) (Seneviratne et al., 2021). 15 16 ENSO is the dominant phenomenon affecting weather conditions in all CSA, and along the Pacific Coast of 17 NWS with effects of heavy rains, storms, floods, landslides, heat and cold waves and extreme sea level rise 18 (Ashok et al., 2007; Reguero et al., 2015; Wang et al., 2017b; Muis et al., 2018; Rodríguez-Morata et al., 19 2018; Rodríguez-Morata et al., 2019; Cai et al., 2020). There is a medium confidence that extreme ENSO 20 will increase long after 1.5°C warming stabilization according to CMIP5 (Cai et al., 2015; Wang et al., 21 2017b; Cai et al., 2018). It is very likely that ENSO rainfall variability, used for defining extreme El Niño 22 and La Niña, will increase significantly, regardless of amplitude changes in ENSO SST variability, by the 23 second half of the 21st century in scenarios SSP2-4.5, SSP3-7.0, and SSP5-8.5 (WGI AR6 Chapter 4; Lee et 24 al., 2021). 25 26 Warming and drier conditions are projected through the reduction of total annual precipitation, extreme 27 precipitation and consecutive wet days, and increase in consecutive dry days (Chou et al., 2014). Heat waves 28 will increase in frequency and severity in places close to the equator as Colombia (Guo et al., 2018; Feron et 29 al., 2019), with decrease but strong wetting in coastal areas, pluvial and river flood, and mean wind increase 30 (Mora et al., 2014). Models project for a 2ºC GWL very likely increase in the intensity and frequency of hot 31 extremes and decrease in the intensity and frequency of cold extremes. Nevertheless, models project 32 inconsistent changes in the region for extreme precipitation (low confidence) (Figure 12.6; WGI AR6 Table 33 12.14; Ranasinghe et al., 2021). The main climate impact drivers in the region, like extreme heat, mean 34 precipitation and coastal and oceanic will increase and snow, ice and permafrost will decrease with high 35 confidence (WGI AR6 Table 12.6, Ranasinghe et al., 2021). 36 37 12.3.2.2 Exposure 38 39 There is high confidence that coastal lowlands are exposed to sea level rise in the form of coastal flooding 40 and erosion, subsidence and saltwater intrusion (Hoyos et al., 2013). Those hazards can affect settlements, 41 ports, industries and other infrastructures. Mangrove and aquaculture areas are among the most exposed 42 systems (Gorman, 2018). The Eastern Tropical Pacific, particularly Sector Niño 3.4, will see the worst 43 increase in sea surface temperature, affecting industrial and small-scale fisheries (very high confidence) 44 (Castrejón and Defeo, 2015; Reguero et al., 2015; Eddy et al., 2019; Bertrand et al., 2020; Castrejón and 45 Charles, 2020; Escobar-Camacho et al., 2021). 46 47 Settlements and agriculture of different scales, and hydroelectric infrastructures, especially near big rivers or 48 in plains, are exposed to floods. Exposure and vulnerabilities to precipitation, overflows and related 49 landslides, are increasing (Briones-Estébanez and Ebecken, 2017). 50 2 In this Report, the following terms have been used to indicate the assessed likelihood of an outcome or a result: Virtually certain 99­100% probability, Very likely 90­100%, Likely 66­100%, About as likely as not 33­66%, Unlikely 0­33%, Very unlikely 0­10%, and Exceptionally unlikely 0­1%. Additional terms (Extremely likely: 95­ 100%, More likely than not >50­100%, and Extremely unlikely 0­5%) may also be used when appropriate. Assessed likelihood is typeset in italics, e.g., very likely). This Report also uses the term `likely range' to indicate that the assessed likelihood of an outcome lies within the 17-83% probability range. Do Not Cite, Quote or Distribute 12-16 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 The Andean piedmont (500­1200 m.a.s.l.) ecosystems and crops and elevation ranges above the treeline are 2 more exposed to thermal anomalies (very high confidence) (Urrutia and Vuille, 2009; Vuille et al., 2015; 3 Aguilar-Lome et al., 2019; Pabón-Caicedo et al., 2020). Temperature rise, combined with precipitation and 4 floods, leave people more exposed to epidemics (very high confidence) (Stewart-Ibarra and Lowe, 2013; 5 Sippy et al., 2019; Petrova et al., 2020). A bigger exposure is related to lower socioeconomic conditions, 6 poor health and marginalisation (Oliver-Smith, 2014). 7 8 12.3.2.3 Vulnerability 9 10 Local economies reliant on limited and specialized resources, highly dependent on ecosystem services such 11 as water and soil fertility, as the alpaca and llama herders or small-scale fishers, are amongst the more 12 vulnerable (very high confidence) (Hollowed et al., 2013; Postigo, 2013; Glynn et al., 2017; Duchicela et al., 13 2019). Also the agricultural sector in the face of extreme events (Coayla and Culqui, 2020). Their 14 vulnerabilities increase as a result of unequal chains of value, incomplete transfers of technology and other 15 socioeconomic and environmental drivers (high confidence) (Ariza-Montobbio and Cuvi, 2020; Gutierrez et 16 al., 2020). 17 18 Informal housing and settlements, usually located in the highest risk land, exacerbates vulnerability (very 19 high confidence) (Miranda Sara and Baud, 2014; Cuvi, 2015; Miranda Sara et al., 2016). The absence of 20 proper drainage systems in urban areas increases the vulnerability, especially to floods. Most of the cities and 21 infrastructure are considered highly vulnerable to climate change (high confidence) (Figure 12.7). 22 23 Regions dependent on glacier runoff are particularly vulnerable (Jiménez Cisneros et al., 2014; Mark et al., 24 2017; Polk et al., 2017). Also biodiversity and water dependent activities where seasonality and rainfall 25 patterns are changing, and where other non-climatic sources of change, such as land use, affect the capacity 26 of ecosystems to provide hydrological services (very high confidence) (Cerrón et al., 2019; Molina et al., 27 2020). The three countries are amongst the most vulnerable in terms of wellbeing and health Figure 12.7; 28 Nagy et al., 2018). 29 30 12.3.2.4 Impacts 31 32 An increase in the frequency of climate related disasters has been reported (high confidence) (Huggel et al., 33 2015a; Stäubli et al., 2018) (WGI AR6 Chapter 12) (Ranasinghe et al., 2021). Scale studies indicate an 34 increase of flood risk during the 21st century, consistent with more frequent floods, being worse in higher 35 emission scenarios (high confidence) (Arnell and Gosling, 2013; Hirabayashi et al., 2013; Alfieri et al., 2017; 36 WGI AR6 Chapter 12, Ranasinghe et al., 2021). Those living on riverbanks and slums built on steep slopes 37 are among the most affected by floods of all kinds (high confidence) (Emmer et al., 2016; Emmer, 2017). 38 There is still uncertainty in relation to future drought intensity and frequency (Pabón-Caicedo et al., 2020). 39 40 Increased sea surface temperature, coupled with stronger ENSO events, will affect marine life and fisheries 41 by loss of productive habitat, disruption of nutrient structure, productivity, and altering the migration of 42 species, leading to changes in fishing rates, impacting coastal livelihoods (high confidence) (Bayer et al., 43 2014; Cai et al., 2015; Ding et al., 2017; Mariano Gutiérrez et al., 2017; Bertrand et al., 2020). Figure 12.8 44 shows other observed sensitivities in several ecosystems and in places as the Galapagos and Malpelo islands, 45 and the coastal Economic Exclusion Zone (EEZ). 46 47 ENSO events coupled with climate change, lead to warmer ocean temperatures, heavy rains, floods and 48 heavy river discharges that have and will impact several activities, including small-scale fisheries 49 infrastructure (very high confidence). In Peru alone, wet extremes are estimated to be at least 1.5 times more 50 likely to happen compared to preindustrial times. The extremely wet ENSO event of 2017 left 6­9 billion 51 USD in monetary losses in that country, 1.7 million inhabitants affected, and crops, roads, bridges, homes, 52 schools, and health posts damaged or destroyed. Distinct types of ENSO events can have differentiated 53 impacts (French and Mechler, 2017; Christidis et al., 2019; Takahashi and Martínez, 2019; Bertrand et al., 54 2020; Coayla and Culqui, 2020). 55 56 Irrigation, potable water, health and education infrastructures, as well as roads, bridges, cities, and housing 57 buildings are frequently damaged or destroyed by extreme precipitations, having also impacts on sediment Do Not Cite, Quote or Distribute 12-17 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 transport, river erosion and annual discharge (very high confidence) (Martínez et al., 2017; Morera et al., 2 2017; Isla, 2018; Rosales-Rueda, 2018; Salazar et al., 2018; Puente-Sotomayor et al., 2021). The increasing 3 variability of precipitation has compromised rain-fed agriculture and power generation, particularly in the 4 dry season (high confidence) (Bradley et al., 2006; Bury et al., 2013; Buytaert et al., 2017; Carey et al., 2017; 5 Vuille et al., 2018; Orlove et al., 2019). For the Amazon-Andes transition zone, impacts of hydrological 6 variability and transport of sediments have been noticed in riparian agriculture and biodiversity (high 7 confidence) (Maeda et al., 2015; Espinoza et al., 2016; Vauchel et al., 2017; Ronchail et al., 2018; Ayes 8 Rivera et al., 2019; Armijos et al., 2020; Figueroa et al., 2020; Pabón-Caicedo et al., 2020). Changes in 9 seasonality and rain patterns are affecting coffee producers (Lambert and Eise, 2020). 10 11 Increases in vector-borne diseases can be related with the increase of rainfall and minimum temperatures 12 during ENSO events (Stewart-Ibarra and Lowe, 2013) and the expansion of the diseases' altitudinal 13 distribution (high confidence) (Lowe et al., 2017; Lippi et al., 2019; Portilla Cabrera and Selvaraj, 2020). 14 ENSO events have been related with diseases such as dengue or leptospirosis (Quintero-Herrera et al., 2015; 15 Sánchez et al., 2017; Arias-Monsalve and Builes-Jaramillo, 2019); they can also increase the incidence of 16 Chikungunya (Section 7.2.2.1; Section 7.3.1.3). Precipitation, relative humidity and temperature have 17 influenced dengue incidence over the last years (Mattar et al., 2013) (Table 12.1). Dengue cases are 18 predicted to increase in the 1.5°C and the 3.7°C warming scenarios by 2050 and 2100, with increases 19 ranging from 28,900 to 88,800 in Peru, 34,600 to 110,000 in Ecuador, and 97,400 to 317,000 in Colombia, 20 although these scenarios do not consider the potential of vaccines or socioeconomic trajectories (Colón- 21 González et al., 2018). Other studies found that Aedes aegypti (arbovirus vector) will shift into higher 22 elevations, increasing the populations at risk (Lippi et al., 2019) (Figure 12.5). Climate change will 23 contribute to increased malaria vectorial capacity (high confidence) (Laporta et al., 2015) (Section 7.2.2.1). 24 Increases in minimum temperature were associated with historical malaria transmission when taking into 25 consideration disease control interventions and climate factors (Fletcher et al., 2020). Figure 12.4 shows 26 mixed changes in the number of months suitable for malaria transmission with low-lying areas in coastal 27 regions becoming more suitable. Zoonotic tick-borne diseases and the epidemiology of tuberculosis are also 28 influenced (Garcia-Solorzano et al., 2019; Rodriguez-Morales et al., 2019). 29 30 31 12-18 Total pages: 181 Do Not Cite, Quote or Distribute FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Figure 12.4: Change in the average number of months in a given year suitable for malaria transmission by Plasmodium 2 falciparum, from 1950-1959 to 2010-2019. The threshold-based model used incorporates precipitation accumulation, 3 average temperature, and relative humidity (Grover-Kopec et al., 2006; Romanello et al., 2021). 4 5 6 Accelerated warming is reducing tropical glaciers. Glacier volume loss and permafrost thawing will continue 7 in all scenarios (high confidence) (Ranasinghe et al., 2021). On average, the tropical Andes have lost about 8 30% and more of their area since the 1980s (Basantes-Serrano et al., 2016; Mark et al., 2017; Thompson et 9 al., 2017; Rabatel et al., 2018; Vuille et al., 2018; Reinthaler et al., 2019a; Seehaus et al., 2019; Masiokas et 10 al., 2020). In a low emissions scenario, by the end of the 21st century, Peru will lose about 50% of the 11 present glacier surface, while in a high-emission scenario there will remain very small areas of only about 3­ 12 5% on the highest peaks (Schauwecker et al., 2017). 13 14 Changing glaciers, snow and permafrost (Figure 12.13), in synergy with land use change, have implications 15 for the occurrence, frequency and magnitude of derived floods and landslides (high confidence) (Huggel et 16 al., 2007; Iribarren Anacona et al., 2015; Emmer, 2017; Mark et al., 2017). Also to landscape transformation 17 through lakes' formation or drying, and to alteration of hydrological dynamics, with impacts on water for 18 human consumption, agriculture, industry, hydroelectric generation, carbon sequestration and biodiversity 19 (high confidence) (Michelutti et al., 2015; Carrivick and Tweed, 2016; Kronenberg et al., 2016; Emmer, 20 2017; Mark et al., 2017; Milner et al., 2017; Polk et al., 2017; Reyer et al., 2017; Young et al., 2017; Vuille 21 et al., 2018; Cuesta et al., 2019; Drenkhan et al., 2019; Hock et al., 2019; Motschmann et al., 2020a). 22 23 Water flow has decreased in several basins as the Shullcas River in the Cordillera Huaytapallana in Peru and 24 is expected to decrease in the near future in in places such as the Cordillera Blanca in Peru (very high 25 confidence) (Baraer et al., 2012; Vuille et al., 2018; Somers et al., 2019; Molina et al., 2020). Disruptions in 26 water flows will significantly degrade or disappear high-elevation wetlands (high confidence) (Bury et al., 27 2013; Dangles et al., 2017; Mark et al., 2017; Polk et al., 2017; Cuesta et al., 2019). Impacts on wetlands are 28 affecting the wild vicuña and the domesticated alpaca (Duchicela et al., 2019). New lakes represent a source 29 of future hazards and water scarcity, as well as an opportunity as water reservoirs (Colonia et al., 2017; 30 Drenkhan et al., 2019). The timing and extent of peak water due to glacier shrinkage is spatially highly 31 variable, and has passed for a large number of tropical Andes glaciers (Hock et al., 2019). Cities dependent 32 on glacier melt have experienced high variability in domestic water supply (Chevallier et al., 2011; Soruco et 33 al., 2015; Mark et al., 2017) as shown in Case Study 2.7.3, but the increase of the demand may also be 34 determinant (Buytaert and De Bièvre, 2012). Water provision is related to socio economic issues (Drenkhan 35 et al., 2015). Glacier retreat impacts Andean pastoralists (high confidence), as shown in Case Study 2.6.5.4. 36 37 NWS houses several global priority areas for biodiversity conservation, including the Tropical Andes and 38 Tumbes-Chocó-Magdalena terrestrial biodiversity-rich spots (Cross-Chapter Paper 1.2.2; Manes et al., 2021) 39 . Biodiversity in Tropical Andes and Tumbes-Chocó-Magdalena is projected to suffer negative impacts 40 (medium confidence: medium evidence, high agreement) (Figure 12.9). Invasive plant species might benefit 41 from climate change in these hotspots (Wang et al., 2017a). Species distribution is changing upslope due to 42 increasing air temperature, leading to range contraction and local extinctions for highland species. Whereas, 43 lowland species are experiencing range contractions at the rear end and expansions in the frontend, including 44 vectors of diseases (high confidence) (Crespo-Pérez et al., 2015; Duque et al., 2015; Morueta-Holme et al., 45 2015; Moret et al., 2016; Aguirre et al., 2017; Cuesta et al., 2017a; Seimon et al., 2017; Fadrique et al., 2018; 46 Tito et al., 2018; Zimmer et al., 2018; Cauvy-Fraunié and Dangles, 2019; Cuesta et al., 2019; Moret et al., 47 2020; Rosero et al., 2021). Vegetation in summits of the northern Andes is particularly vulnerable because of 48 a high abundance of endemic species with narrow thermal niches, and lowland dispersal capacity in 49 comparison to the Central Andes (Cuesta et al., 2020). 50 51 The upper limit of alpine vegetation (paramo) shifted upslope 500 m in the Chimborazo (Morueta-Holme et 52 al., 2015). Yet, the upper forest limit (the ecotone between forest and alpine vegetation), is migrating at 53 slower rates, or not migrating at all (Harsch et al., 2009; Rehm and Feeley, 2015b), so it is expected to be a 54 major barrier to migration to several montane species, leading to population reductions and biodiversity 55 losses (Lutz et al., 2013; Rehm and Feeley, 2015a). Shifts in tree species distribution may result in decreased 56 above ground carbon stocks and productivity in tropical mountain forests (high confidence) (Feeley et al., 57 2011; Duque et al., 2015; Fadrique et al., 2018; Duque et al., 2021), a biomass loss that will only be partially Do Not Cite, Quote or Distribute 12-19 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 offset through increased recruitment and growth of lowland species migrating upslope. Water scarcity can 2 enhance tree mortality and decrease above ground carbon stocks (Álvarez-Dávila et al., 2017; McDowell et 3 al., 2020). Agricultural frontier of crops, as potatoes or maize, is going upwards (high confidence), following 4 the freezing level height upward displacement (Morueta-Holme et al., 2015; Skarbø and VanderMolen, 5 2016; Schauwecker et al., 2017; Vuille et al., 2018). Modelling exercises agree with the observed impacts in 6 species, ecosystem processes, crop impacts and related pests and diseases (high confidence) (Cernusak et al., 7 2013; Tovar et al., 2013; Ramirez-Villegas et al., 2014; Ovalle-Rivera et al., 2015; van der Sleen et al., 2015; 8 Lowe et al., 2017). Agricultural options are changing as a result of intra seasonal temperature variation 9 (Ponce, 2020). Changes in timing and amount of precipitation are also impacting agriculture (Table 12.4; 10 Heikkinen, 2017; Altea, 2020) . 11 12 Species distribution is changing in dry lowland forests, where deforestation is the more intense driver and 13 climate change is intensely acting (Aguirre et al., 2017; Manchego et al., 2017). Extinctions in amphibians 14 have been related with temperature raises acting in synergy with diseases (Catenazzi et al., 2014). The 15 fungus Batrachochytrium dendrobatidis successfully accompanied and caused disease in high-elevation 16 Andean frogs as they expanded their ranges to reach 5200­5400 m (Seimon et al., 2017). Several groups of 17 freshwater species of the tropical Andes represent 35% of threatened freshwater species in the world 18 (Gardner and Finlayson, 2018). Potential impacts of species turnover in key areas for biodiversity 19 conservation have been identified (Cuesta et al., 2017b). 20 21 Climate change related hazards could foster rural poverty, and its impacts have led to the modification of 22 agriculture calendars and irrigation adjustments (Postigo, 2014). Livestock is reducing due to rising 23 temperatures, changing water flows and diminishing of pastures, particularly cattle and pig production 24 (Bayer et al., 2014; Tapasco et al., 2015; Bergmann et al., 2021). In some cases farmers respond to extreme 25 temperatures by increasing use of land and crop intensity (Aragón et al., 2021). Climate change has and will 26 prompt internal and international migrations (Løken, 2019; Bergmann et al., 2021). A change in fire regimes 27 and fire risk is expected in highland ecosystems, although it is difficult to determine the influence of human 28 activities and climate change influence on fire patterns (Oliveras et al., 2014; Oliveras et al., 2018; 29 Armenteras et al., 2020). 30 31 12.3.3 Northern South America (NSA) Sub-region 32 33 12.3.3.1 Hazards 34 35 A significant increase in the intensity and frequency of warm extremes and length of heat waves, and 36 decrease in the frequency of cold extremes (Skansi et al., 2013) was likely observed (Figure 12.6; Donat et 37 al., 2013; Almeida et al., 2017; WGI AR6 Table 11.13, Seneviratne et al., 2021). Precipitation showed 38 increasing trends in annual and wet season totals over the eastern part and decreasing trends of the dry 39 season (Almeida et al., 2017). Increase in the frequency of anomalous severe floods (Gloor et al., 2015) was 40 observed but insufficient data coverage for extreme precipitation and trends in available data result in low 41 confidence (Avila-Diaz et al., 2020; Dereczynski et al., 2020; Dunn et al., 2020; Sun et al., 2021) (WGI AR6 42 Table 11.14) (Seneviratne et al., 2021). Droughts presented mixed trends between subregions, but evidences 43 indicate increasing length of dry periods (low confidence) (Skansi et al., 2013; Marengo and Espinoza, 2016; 44 Spinoni et al., 2019; Avila-Diaz et al., 2020; Dereczynski et al., 2020; Dunn et al., 2020) (WGI AR6 Table 45 11.15) (Seneviratne et al., 2021) (WGI AR6 Table 12.3) (Ranasinghe et al., 2021). 46 47 An overall increase in temperature by the end of century is projected for all the seasons, from 2 to 6°C 48 depending on the scenario (Chou et al., 2014). Projections also suggest increases in the intensity and 49 frequency of hot extremes and decreases in the intensity and frequency of cold extremes (very likely for a 50 2ºC GWL) (López-Franca et al., 2016) (WGI AR6 Table 11.13) (Seneviratne et al., 2021). In all the region, 51 extreme maximum temperature estimates under the RCP4.5 scenario are projected to increase. Tropical 52 major cities are expected to be strongly affected by heat waves and daily record temperatures (Feron et al., 53 2019). 54 55 A decrease in precipitation over the tropical region but regional changes, such as increases in rainfall 56 amounts in western NSA of up to 40 mm, are expected by mid-century under RCP8.5 (Teichmann et al., 57 2013; Sánchez et al., 2015). Changes in the dry season in the central part of South America due to the late Do Not Cite, Quote or Distribute 12-20 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 onset and late retreat of monsoon, decreases in precipitation over the Amazon and central Brazil are expected 2 (Coppola et al., 2014; Giorgi et al., 2014; Llopart et al., 2014). And with medium confidence: increase in the 3 frequency and geographic extent of meteorological drought in the eastern Amazon, and the opposite in the 4 west (Duffy et al., 2015). A decreasing of total annual precipitation, but increase in heavy precipitation 5 (Seiler et al., 2013; Chou et al., 2014) are projected for a 2ºC GWL (Figure 12.6; WGI AR6 Table 11.15; 6 Seneviratne et al., 2021). 7 8 Mean precipitation will decrease and heavy precipitation, aridity and drought will increase with medium 9 confidence, mean temperature, extreme heat, fire weather, coastal and oceanic climate impact drivers all of 10 them will increase with high confidence (Sun et al., 2019) (WGI AR6 Table 12.6; WGI AR6 Figure 12.8) 11 (Ranasinghe et al., 2021). 12 13 12.3.3.2 Exposure 14 15 In NSA the percentage of the national population living in Low Elevation Coastal Zones (LECZ) and 16 exposed to Sea Level Rise (SLR) is 68% for Suriname, 56% for Guyana and 6% Venezuela (Nagy et al., 17 2019). In these countries, exposure of populations, land areas and built capital to coastal floods is projected 18 to continue and increase (Neumann et al., 2015; Reguero et al., 2015). 19 20 In the Amazon basin, approximately 80% of the population is concentrated in cities due to migrations in 21 search of improvements in education, job opportunities, health and goods and services (Eloy et al., 2015; 22 Pinho et al., 2015). These populations settle in areas prone to flooding combined with various levels of 23 sanitation due to limited economic access to areas of lower risk (Pinho et al., 2015; Mansur et al., 2016; 24 Andrade and Szlafsztein, 2018; Parry et al., 2018). In these areas, poor urban planning and high population 25 densities increase exposure levels (Mansur et al., 2016). In this context, 41% of the total population of urban 26 centres, of the Amazon Delta and Estuaries (ADE) are exposed to flooding (Mansur et al., 2016), while in 27 Santarem, population and infrastructure are highly exposed to floods and flash floods (Andrade and 28 Szlafsztein, 2018). 29 30 Exposure of the Brazilian Amazon to severe to extreme drought has increased from 8% in 2004/2005, to 31 16% in 2009/2010 and 16% in 2015/2016 (Anderson et al., 2018b); a similar trend is reported in other 32 regions (Table 12.3). During the extreme drought of 2015/2016 in the Amazonian forests 10% or more of the 33 area showed negative anomalies of the Minimum Cumulative Water Deficit (Anderson et al., 2018b). This 34 extreme drought also caused an increase in the occurrence and spread of fires in the basin (medium 35 confidence: medium evidence, high agreement) (Aragão et al., 2018; Lima et al., 2018; Silva Junior et al., 36 2019; Bilbao et al., 2020).The exposure to anomalous fires in ecosystems such as savannas, more fire-prone, 37 increases the exposure and vulnerability of adjacent forest ecosystems not adapted to fire, such as seasonally 38 flooded forests (Bilbao et al., 2020; Flores and Holmgren, 2021). 39 40 12.3.3.3 Vulnerability 41 42 NSA is one of the most vulnerable subregions in the region, after CA, as evidenced by its very high 43 vulnerability in four of the six sectors assessed (Figure 12.7). LECZ of Venezuela, Guyana and Suriname are 44 highly vulnerable to climate change due to SLR (high confidence) (CAF, 2014; Mycoo, 2014; Reguero et al., 45 2015; Villamizar et al., 2017; Nagy et al., 2019). In Guyana, the combined effect of increased rainfall 46 intensity and SLR has caused flooding over the past two decades, increasing the vulnerability of the 47 agriculture sector (Tomby and Zhang, 2019). 48 49 The unprecedented extreme events of floods (2009, 2012 and 2014) and drought (2010) in the Amazon basin 50 led to increased societies vulnerability (medium confidence: medium evidence, high agreement) (Mansur et 51 al., 2016; Debortoli et al., 2017; Marengo et al., 2018; Menezes et al., 2018). The disruption of the region 52 natural hydrology dynamics, as a consequence of extreme events increases the sensitivity of the food and 53 transport systems of the Indigenous Peoples and rural resource-dependent communities (Pinho et al., 2015). 54 55 Migration by Indigenous Peoples and rural resource-dependent communities to cities have increased due to 56 urbanization, development of extractive activities, agroindustry and infrastructure. Upon migrating, they are 57 forced to abandon their livelihoods in order to acquire temporary jobs and to live in poverty and exclusion Do Not Cite, Quote or Distribute 12-21 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 conditions on the periphery of the city (Cardoso et al., 2018). Between 60­90% of the population in the 2 urban centres of ADE live in conditions of moderate to high degree of vulnerability (Mansur et al., 2016) 3 (Figure 12.7). Amazon populations located in remote urban centres with limited or non-existing roads are 4 more vulnerable to extreme events in relation to more connected urban centres (Parry et al., 2018). These 5 highly vulnerable circumstances reduce the adaptive capacity of these populations (Cardoso et al., 2018). 6 Nevertheless, the dynamics of the adaptive capacity of the Indigenous Peoples and rural resource-dependent 7 communities is a complex issue. There is robust and growing literature showing that resource-dependent 8 communities located in remote areas, address climate anomalies by reducing the vulnerability of socio- 9 ecological systems through Indigenous knowledge and local knowledge (high confidence) (Mistry et al., 10 2016; Vogt et al., 2016; Bilbao et al., 2019; Bilbao et al., 2020; Camico et al., 2021). 11 12 Amazonian forests constitute one of the major carbon (C) sinks on Earth (Pan et al., 2011), playing a pivotal 13 role in the climate system and regional balance of C and water (Marengo et al., 2018; Molina et al., 2019). 14 Deforestation, temperature increase and any factor affecting the forests ecosystem dynamics will have an 15 impact on the atmospheric CO2 concentration and hence on the global climate (Ruiz-Vásquez et al., 2020; 16 Sullivan et al., 2020).There is robust scientific evidence of the high vulnerability of the Amazonian forests 17 to increasing temperature and repeated extreme drought events (high confidence) (Figure 12.7; Brienen et al., 18 2015; Olivares et al., 2015; Feldpausch et al., 2016; Zhao et al., 2017; Anderson et al., 2018b; Anjos and De 19 Toledo, 2018; Yang et al., 2018; Barkhordarian et al., 2019; Sampaio et al., 2019; Rammig, 2020; Sullivan et 20 al., 2020) . 21 22 12.3.3.4 Impacts 23 24 Suriname has experienced coastal erosion and flooding, causing damage to infrastructure, agriculture and 25 ecosystems while Georgetown has suffered a significant number of floods (CAF, 2014). In Guyana, coastal 26 flooding has negatively impacted agricultural activity (Tomby and Zhang, 2018) (Figure 12.9). Sugarcane 27 production has been one of the most impacted cash-crops. The impact on sugar production has affected 28 Guyana's sugar industry (Tomby and Zhang, 2019). Among the main impacts observed in the sugar industry 29 are an increase in production costs, greater use of pesticides and fertilizers, and a reduction in workers' 30 income (Tomby and Zhang, 2018). 31 32 Indigenous Peoples and resource-dependent rural communities in the Amazon have been impacted over the 33 last decade by extreme drought and flood events in various dimensions of their livelihoods (Pinho et al., 34 2015). Food security has been strongly impacted since it is based on fishing and small-scale agriculture, two 35 sectors highly vulnerable to climate change. During extreme events, fishing decreases due to limited access 36 to fishing grounds (medium confidence: low evidence, high agreement) (Figure 12.9; Pinho et al., 2015; 37 Camacho Guerreiro et al., 2016. Overfishing, deforestation and dam construction are a threat to fishing in the 38 subregion (Lopes et al., 2019) and therefore contribute to exacerbating the impacts of climate change. Small 39 scale agriculture practices (e.g., floodplain agriculture and slash and burn), are highly coupled with natural 40 hydrological cycles and therefore severely affected by extreme events (Figure 12.9; Cochran et al., 2016. 41 Livelihoods are also impacted by disruptions in land and river transport, restrictions in drinking water access, 42 increased incidence of forest fires and disease outbreaks (medium confidence: medium evidence, high 43 agreement) (Figure 12.9; Marengo et al., 2013; Pinho et al., 2015; Marengo and Espinoza, 2016; Marengo et 44 al., 2018). In addition, flood events have caused losses of homes and disruption of public and commercial 45 services (Figure 12.9; Parry et al., 2018). 46 47 Several vector-driven diseases such as malaria and leishmaniasis are endemic of Amazon region, however 48 socio-environmental changes are altering their natural dynamics (Confalonieri et al., 2014b). An important 49 relationship between the outbreak of infectious diseases and changes in climatic events (e.g., droughts, 50 floods, heat waves, ENSO) or environmental events (e.g., deforestation, dam construction and habitat 51 fragmentation) have been found for the Brazilian Amazon (medium confidence: medium evidence, high 52 agreement) (Pan et al., 2014; Filho et al., 2016; Nava et al., 2017; Ellwanger et al., 2020). These impacts are 53 more severe in poor populations with limited access to health services (Pan et al., 2014; WHO and 54 UNFCCC, 2020). In the case of Venezuela, the impact of climate change on the epidemiology of malaria has 55 been studied, showing significant influence on transmission in the Amazonia area of the country (Figure 56 12.4; Laguna et al., 2017) . Other studies from Venezuela have documented the role of ENSO in dengue 57 outbreaks (Vincenti-Gonzalez et al., 2018). Table 12.1 shows the changes observed in reproduction potential Do Not Cite, Quote or Distribute 12-22 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 for dengue in the different subregions due to changes in rainfall and temperature. Forest fires are a major 2 concern to public health in the region as they relate to an increase in hospital admissions due to respiratory 3 problems, mainly among children and the elderly (Figure 12.5). The amount of air pollutants detected is 4 sometimes higher than that observed in large urban areas, especially during dry seasons when biomass 5 burning increases (Aragão et al., 2016; de Oliveira Alves et al., 2017; Paralovo et al., 2019). 6 7 8 Table 12.1: Environmental suitability for the transmission of dengue by Aedes aegypti as modelled by the influence of 9 temperature and rainfall on vectorial capacity and vector abundance; this is overlaid with human population density data 10 to estimate the reproduction potential for these diseases (R0, the expected number of secondary infections resulting from 11 one infected person). The Southwest South America (SWS) and Southern South America (SSA) subregions are not 12 presented, as the vector is not abundant in these areas and the estimated R0 is lower than 0.01. Data derived from 13 Romanello et al. (2021). Subregion Average R0 Average R0 Absolute change in % change in R0 1950-1954 2016-2020 R0 from 1950-54 to from 1950-54 to 2016-20 2016-21 Central America (CA) 3.00 3.53 0.53 18% Northwest South America (NWS) 1.85 2.40 0.55 30% Northern South America (NSA) 1.31 2.05 0.74 56% South America Monsoon (SAM) 0.93 1.67 0.74 80% Northeast South America (NES) 2.11 2.47 0.36 17% Southeast South America (SES) 0.64 0.81 0.17 26% 14 15 16 Climate change impacts have also been observed in ocean, coastal ecosystems (coral reefs and mangroves), 17 Exclusive Economic Zones (EEZ) and saltmarshes in NSA; further impacts are expected in coral reefs, 18 estuaries, mangroves and EEZs in the sub-region (Figure 12.9). Species in freshwater ecoregions (e.g., the 19 Orinoco and Amazon Rivers and their flooded forests) are predicted to suffer a decrease in range and 20 climatic suitability (medium confidence: low evidence, high agreement) (Cross-Chapter Paper 1.2.3; Manes 21 et al., 2021) . A significant decrease in climate refugia (90%) for multiple vertebrate and plant species in the 22 region has been projected for a 4ºC scenario, with considerable benefits of mitigation and reducing risks to 23 40% for a 2oC scenario (Warren et al., 2018). 24 25 Droughts in 2009/2010 and 2015/2016 increased tree mortality rate in Amazon forests (Doughty et al., 2015; 26 Feldpausch et al., 2016; Anderson et al., 2018b), while productivity didn't show a consistent change; some 27 authors report a drop in productivity (Feldpausch et al., 2016) and others found no significant changes 28 (Brienen et al., 2015; Doughty et al., 2015). Nevertheless the combined effect of increasing tree mortality 29 with variations in growth, results in a long-term decrease in C stocks in forest biomass compromising their 30 role of these forests as C sink (high confidence) (Brienen et al., 2015; Rammig, 2020; Sullivan et al., 2020) 31 (Figure 12.9). Under the RCP8.5 scenario for 2070, drought will increase the conversion of rainforest to 32 savannahs (medium confidence: medium evidence, high agreement) (Anadón et al., 2014; Olivares et al., 33 2015; Sampaio et al., 2019). The transformation of rainforest into savannahs brings forth biodiversity loss 34 and alterations in ecosystem functions and services (medium confidence: medium evidence, high agreement) 35 (Anadón et al., 2014; Olivares et al., 2015; Sampaio et al., 2019). In the Amazon basin, the synergic effects 36 of deforestation, fire, expansion of the agricultural frontier, infrastructure development, extractive activities, 37 climate change and extreme events may exacerbate the risk of savannisation (medium confidence: medium 38 evidence, high agreement) (Nobre et al., 2016b; Bebbington et al., 2019; Sampaio et al., 2019; Rammig, 39 2020). 40 41 12.3.4 South America Monsoon (SAM) Sub-region 42 43 12.3.4.1 Hazards 44 Do Not Cite, Quote or Distribute 12-23 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Temperature extremes have likely increased in the intensity and frequency of hot extremes and decrease in 2 the intensity and frequency of cold extremes (Donat et al., 2013; Bitencourt et al., 2016) (WGI AR6 Table 3 11.13) (Seneviratne et al., 2021). In a vast transition zone between the Amazon and the Cerrado Biomes 4 within the region, analysis of seasonal precipitation trends suggested that almost 90% of the observational 5 sites showed reduced in the length of the rainy season in the region (Debortoli et al., 2015), on a period from 6 1971 to 2014 (Marengo et al., 2018), confirming the growth in length of the dry season. Changes in the 7 hydrological and precipitation regimes, characterized by reduction in rainfall in Southern Amazonia, 8 contrasting to an increase in the northwest Amazonia, and overall increases in extreme precipitation and in 9 the frequency of Consecutive Dry Days, is being reported by several authors (Fu et al., 2013; Almeida et al., 10 2017; Marengo et al., 2018; Espinoza et al., 2019a) with low confidence (WGI AR6 Table 11.14; 11 Seneviratne et al., 2021) due to insufficient data coverage and trends in available data generally not 12 significant. 13 14 The Amazon has been identified as one of the areas of persistent and emergent regional climate change 15 hotspots in response to various representative concentration pathways (Diffenbaugh and Giorgi, 2012). In 16 Bolivia, CMIP3/5 models projected an increase in temperature (2.5ºC­5.9ºC), with seasonal and regional 17 differences. In the lowlands, both ensembles agreed on less rainfall (­19%) during drier months (June­ 18 August and September­November), with significant changes in inter-annual rainfall variability, but 19 disagreed on changes during wetter months (January­March) (Seiler et al., 2013). As a consequence of 20 higher temperatures and reduced rainfall, an increased water deficit would be expected in the Brazilian 21 Pantanal (Marengo et al., 2016; Bergier et al., 2018; Llopart et al., 2020) with high confidence. The largest 22 increases in warmer days and nights, and aridity, drought and significant increases in fire occurrence are 23 calculated over the Amazon area (Huang et al., 2016). Over all the region, by mid-century (RCP4.5) there is 24 medium confidence of increase of river and pluvial floods, aridity and mean wind speed, and extreme heat, 25 fire weather and drought are projected to increase with high confidence (WGI AR6 Table 12.6; Ranasinghe 26 et al., 2021). 27 28 12.3.4.2 Exposure 29 30 A large expansion in cropland area (soybean, corn and sugarcane) was observed in the past two decades in 31 SAM, in response to an increased local and global demand for biofuels and agricultural commodities (high 32 confidence) (Lapola et al., 2014; Cohn et al., 2016). Feedbacks to the climate system resulting from such 33 land-use changes are intricate. The clear-cutting of Amazon forest and Cerrado savannah in the region lead 34 to a local warming due to an increase in the energy balance and evapotranspiration (Malhado et al., 2010), 35 contrastingly the replacement of pasture by agriculture leads to local cooling effect, due to changes in the 36 surface albedo (medium confidence: medium evidence, medium agreement). Deforestation of the Amazon for 37 pastures and soybean have decreased evapotranspiration during drought months and caused a localized 38 lengthening of the dry season in Northwest SAM by 6.5 (± 2.5) days since 1979 (medium confidence: 39 medium evidence, medium agreement) (Fu et al., 2013). 40 41 It is not surprising therefore that while SAM is the region in CSA that experienced the highest temperature 42 increase in the last century, it is where most of the fire spots in the sub-continent are located, owing also to 43 the prevalent use of fires in pasture lands (medium confidence: medium evidence, high agreement) (Bowman 44 et al., 2009). Recently, da Silva Junior et al. (2020) reported 6,708,350 and 6,188,606 fire foci in Cerrado 45 and Amazonia, between 1999 and 2018, corresponding to 80% of the total observed in Brazil. The 46 occurrence of extreme droughts has affected the carbon and water cycles in large areas of the Amazon Forest 47 (high confidence) (Lapola et al., 2014; Agudelo et al., 2019), in particular in its southern and eastern 48 portions, where deforestation rates are higher. The loss of carbon in the Amazon region considering the 49 combined effect of land use change in the southern portion of the region, bordering Cerrado and Pantanal, 50 and global carbon emission scenarios, can be up to 38% at 4ºC of warming, but limited to 8% if the Paris 51 agreed limit of 1.5°C is achieved (medium confidence, medium evidence, high agreement) (Burton et al., 52 2021), driving the region to be a net carbon source to the atmosphere (Gatti et al., 2021). A recent extreme 53 drought was estimated to affect the photosynthetic capacity of 400,000 km2 of the forest (Anderson et al., 54 2018b), nevertheless there are considerable uncertainties regarding the effects of CO2 fertilization in tropical 55 forests and ecosystems (medium confidence: medium evidence, high agreement) (Sampaio et al., 2021). 56 Extreme drought events increase forest vulnerability to fire, directly affecting the biodiversity, the forest 57 structure and its plant species distribution (high agreement) (Brando et al., 2014). Production sectors are also Do Not Cite, Quote or Distribute 12-24 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 exposed. SAM is pointed out as a region where agricultural production will be especially impacted by 2 climate change, affecting production of annual crops, fruits and livestock (medium confidence: medium 3 evidence, high agreement) (Lapola et al., 2014; Zilli et al., 2020). 4 5 12.3.4.3 Vulnerability 6 7 The largest expanses of remaining vegetation in the Cerrado biome are located in SAM, but the region shows 8 low number of protected areas (only 7.5% inside protected areas), which will leave fauna and flora with little 9 room for moving across the landscape in the face of climate change. Protected areas --Indigenous lands 10 included-- have markedly detained forest clear-cutting in the Amazon deforestation arc (most of which is 11 inside SAM) (high confidence) (Nolte et al., 2013). However nearly one hundred protected areas in the 12 Amazon, Cerrado and Pantanal biomes inside SAM have been identified as highly or moderately vulnerable 13 to future climate change and demand deep adaptation interventions (medium confidence: medium evidence, 14 high agreement) (Feeley and Silman, 2016; Lapola et al., 2019b). Yet, the maintenance of these protected 15 areas or even the halting of deforestation may do little to impede a large-scale ecosystem shift, persistently, 16 to an alternative state (crossing a tipping point) of the Amazon forest or even more subtle changes caused by 17 climate change in the region (medium confidence: medium evidence, high agreement) (Aguiar et al., 2016a; 18 Boers et al., 2017; Lapola et al., 2018; Lovejoy and Nobre, 2018). 19 20 The agriculture in the region is highly dependent on the climate (high confidence), responsible for ¾ of the 21 variability in agricultural yields in the region (Table 12.4). Irrigation is an important strategy for agriculture 22 production in part of the region, nevertheless not accounting to more than 8% of the total agricultural area in 23 South America and 7% in Central America (OECD and FAO, 2019). This practice faces potential impacts 24 from reduction in surface water availability in future climate scenarios (Ribeiro Neto et al., 2016; Zilli et al., 25 2020), enhanced by non-climate drivers such as land use changes (medium confidence: medium evidence, 26 high agreement) (Spera et al., 2020). The remaining fluctuation on yields relates to issues of infrastructure, 27 market, economy, policy and social aspects. Good infrastructure, transport logistics, quality of roads and 28 storage, strongly influences the vulnerability of the agriculture sector (Figure 12.7). 29 30 The combined effect of extreme climate events and ecosystem fragmentation, e.g., by deforestation or fire, 31 lead to changes in forest structure, with the death of taller trees and reduction in diversity of plant species, 32 loss of productivity and carbon storage (high agreement) (Brando et al., 2014; Reis et al., 2018). The rise of 33 the large-scale soybean agroindustry in the early 2000's led to a faster increase in human development 34 indicators in some regions, tightly linked to the agricultural production chain (high confidence) (Richards et 35 al., 2015). Such a development also came at a considerable cost for the environment (e.g., Neill et al. (2013)) 36 and the regional climate, even though a moratorium implemented in 2006 to refrain new soy plantations on 37 deforested areas reduced deforestation by a factor of five (high confidence) (Macedo et al., 2012; Kastens et 38 al., 2017). The same sort of supply chain interventions along with incentive-based public policies applied to 39 the beef supply chain could minimize the need for agricultural expansion in the SAM deforestation frontier 40 (medium confidence: medium evidence, high agreement) (Nepstad et al., 2014; Pompeu et al., 2021). 41 42 SAM has a low population density, and the majority of population is located in cities. The population of 43 some of these cities are indicated as highly vulnerable considering the enormous social inequalities 44 embedded in these cities (high confidence) (Filho et al., 2016). Inequalities and uneven access to 45 infrastructure, housing and health support, increase population vulnerability to atmospheric pollution and 46 drier conditions (high confidence) (Rodrigues et al., 2019; IPAM, 2020; Machado-Silva et al., 2020). 47 48 12.3.4.4 Impacts 49 50 The Amazon and the Cerrado are amongst the largest and unique phytogeographical domains in South 51 America. The Brazilian Cerrado is amongst the richest biodiversity in the world, with more than 12,600 plant 52 species, being 35% endemic (high confidence) (Forzza et al., 2012). Historic land cover change and 53 concurrent climate change in the region strongly impacted the biodiversity and led to the extinction of 657 54 plant species for the Cerrado, which is more than four-fold the global recorded plant extinctions (high 55 confidence) (Strassburg et al., 2017; Green et al., 2019). Effects of climate change, expressed by drought and 56 heat waves, lead to plant stress, compromising growth and increasing mortality (Yu et al., 2019). The fauna 57 dependent on dew water was strongly impacted due to a temperature rise of 1.6ºC from 1961 to 2019 Do Not Cite, Quote or Distribute 12-25 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 (medium confidence: medium evidence, medium agreement) (Hofmann et al., 2021). Modelling outcomes 2 project impacts in forest ecosystems in the region, with persistent warming and significant moisture 3 reduction (Anjos et al., 2021), leading to a potential change in the ecosystem structure and distribution in the 4 region (medium confidence: medium evidence, medium agreement) (Government of Brazil, 2020). 5 6 The observed impact on plant species in SAM is projected to worsen in a warmer world (Warszawski et al., 7 2013). An increasing dominance of drought-affiliated genera of tree species has been reported in the 8 southern part of the Amazon forest in the last 30 years (medium confidence: medium evidence, medium 9 agreement) (Esquivel-Muelbert et al., 2019). Due to the tight relation of drought and fire occurrence, an 10 increase of 39 to 95% of burned area is modelled to impact the Cerrado region under RCP4.5 and RCP8.5, 11 while under RCP2.6, a 22% overshoot in temperature is estimated to impact the area in 2050 decreasing to 12 11% overshoot by 2100 (Silva et al., 2019d), leading to high impact on agriculture production (high 13 confidence). 14 15 SAM hosts the headwaters of important SA rivers such as the Paraguay, Madeira, Tocantins-Araguaia and 16 Xingu. The impact from climate change is expressed differently among several sub-regions. Extreme floods 17 in Southern Amazon and Bolivian Amazon floodplains were described and related to exceptionally warm 18 subtropical South Atlantic ocean (high confidence) (Espinoza et al., 2014), causing high economic impact 19 (losses in crop and livestock production and infrastructure) and number of fatalities (very high confidence) 20 (Ovando et al., 2016). Contrastingly, decline in stream flow, particularly in the dry season, expressed by the 21 ratio between runoff and rainfall, is observed for the southern part of the Amazon basin (high evidence) 22 (Molina-Carpio et al., 2017; Espinoza et al., 2019b; Heerspink et al., 2020). Observed precipitation reduction 23 in the Cerrado region impacted main water supply reservoir for important cities in the Brazilian central 24 region, leading to a water crisis in 2016/2017 (Government of Brazil, 2020) and affecting energy 25 hydropower generation (Ribeiro Neto et al., 2016). Modelling studies project decreases in river discharge 26 rate in the order of 27% for the Tapajós basin and 53% for the Tocantins-Araguaia basin for the end of the 27 century, which may affect freshwater biodiversity, navigation and generation of hydroelectric power 28 (medium confidence: medium evidence, high agreement) (Marcovitch et al., 2010; Mohor et al., 2015). This 29 region also holds one of the largest floodplains in the globe, the Pantanal. The climatic connection of 30 Pantanal regions to the Amazon, and the influence of deforestation in local precipitation (Marengo et al., 31 2018) has implications for conservation of ecosystem services and water security in Pantanal (high 32 confidence) (Bergier et al., 2018). Impacts of extreme drought, with increasing numbers of dry days, and 33 peak of fire foci was recently reported (robust evidence) (Lázaro et al., 2020; Garcia et al., 2021). Projected 34 impacts of climate change shall lead to profound changes in the annual flood dynamics for the Pantanal 35 wetland, altering ecosystem functioning and severely affecting biodiversity (high confidence) (Thielen et al., 36 2020; Marengo et al., 2021) 37 38 Soybean and corn yields, in the Cerrado region, will suffer one of the strongest negative impacts under 39 RCP4.5 and RCP8.5 scenarios estimate and will demand high investments for adaptation should it continue 40 to be cultivated in the same localities as today (high confidence) (Oliveira et al., 2013; Camilo et al., 2018). 41 Changes in precipitation patterns were related to reduction of agriculture productivity and revenues in the 42 southern portion of the Amazon region (medium confidence: medium evidence, high agreement) (Costa et 43 al., 2019; Leite-Filho et al., 2021). As such, the future socio-economic vigour of the region will be, to a large 44 extent, connected to an unlikely stability of the regional climate and eventual fluctuations of global markets 45 potentially affecting the agricultural supply chain (high confidence) (Nepstad et al., 2014). 46 47 Observations from recent past droughts in SAM indicates how the incidence of respiratory diseases may 48 worsen under a drier and warmer climate. Northwest SAM had a ~54% increase in the incidence of 49 respiratory diseases associated with forest fires during the 2005 drought compared to a no-drought 10-year 50 mean (high confidence) (Ignotti et al., 2010; Pereira et al., 2011; Smith et al., 2014). It is estimated that more 51 than 10 million people are exposed to forest fires in the deforestation arc, a region comprising several 52 Brazilian states in the southern and western parts of the Amazon forest, with several impacts on human 53 health including potential exacerbation the COVID-19 crisis in Amazonia (medium confidence: medium 54 evidence, high agreement) (de Oliveira et al., 2020) (Table 12.5). Increases in hospital admissions, asthma, 55 DNA damage and lung cell death due to inhalation of fine particulate matter, represents an increase in public 56 health system costs (high confidence) (Ignotti et al., 2010; Silva et al., 2013; de Oliveira Alves et al., 2017; 57 Machin et al., 2019).The patchy landscape created by forest clearing contribute to a rising risk of zoonotic Do Not Cite, Quote or Distribute 12-26 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 disease emergence by increasing interactions between wildlife, livestock and humans (medium confidence: 2 low evidence, medium agreement) (Dobson et al., 2020; Tollefson, 2020). Recent studies also suggested the 3 influence of climate change in zoonotic diseases, such as Orthohantavirus and Chapare virus infections, 4 rodent-borne diseases, in some areas of Bolivia (Escalera-Antezana et al., 2020a; Escalera-Antezana et al., 5 2020b). Extreme fluctuation in the river level in the amazon was associated to a significant increase in the 6 incidence of diarrhoea, leptospirosis and dermatitis (de Souza Hacon et al., 2019; Government of Brazil, 7 2020). A comprehensive characterization of future heatwaves, and alternative RCPs scenarios, Brazilian 8 urban areas at SAM region are projected to face increasing related mortality from 400 to 500% in the period 9 from 2031 to 2080 compared to the period of 1971­2020, under the highest emission scenario and high- 10 variant population scenario (medium confidence: low evidence, medium agreement) (Guo et al., 2018). Table 11 12.2 shows the increase in days of exposure to heatwaves already observed in the region. 12 13 14 Table 12.2: Average change in the mean number of days exposed to heatwaves (defined as a period of at least two days 15 where both the daily minimum and maximum temperatures are above the 95th percentile of their respective climatologies) 16 in the population over 65 years of age in 2016-2020 relative to 1986-2005. Temperature data taken from the European 17 Centre for Medium-Range Weather Forecasts (ECMWF) ERA5 dataset; calculations derived from Romanello et al. 18 (2021). Country Number of additional days of heatwave exposure in 2016- 2020 relative to 1986-2005 Argentina 4.9 Belize 8.8 Bolivia 2.2 Brazil 3.1 Chile 3.3 Colombia 9.3 Costa Rica 0.8 Ecuador 7.6 El Salvador 2.2 Guatemala 8.4 Guyana 8.2 Honduras 11.2 Nicaragua 2.2 Panama 2.6 Paraguay 2.6 Peru 3.6 Suriname 15.2 Uruguay 2.7 Venezuela 8.5 19 20 21 The high risk of floods (high-frequency and high-incurred damage) is centred in the Brazilian states of Acre, 22 Rondônia, Southern Amazonas and Pará (Andrade et al., 2017). Global-scale studies indicate an increase of Do Not Cite, Quote or Distribute 12-27 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 flood risk for the SAM region during the 21st century (consistent with floods that are more frequent) (high 2 confidence) (Hirabayashi et al., 2013; Arnell et al., 2016; Alfieri et al., 2017). Higher emission scenarios 3 result in substantially higher flood risks than low emission scenarios (Alfieri et al., 2017). 4 5 12.3.5 Northeast South America (NES) Sub-region 6 7 12.3.5.1 Hazards 8 9 The region has likely experienced an increase in temperature, with significant increases in the intensity and 10 frequency of hot extremes and significant decreases in the intensity and frequency of cold extremes (Donat et 11 al., 2013) (WGI AR6 Table 11.13, Seneviratne et al., 2021). 12 13 A decrease in the frequency and magnitude of extreme precipitation was observed but with low confidence, 14 due to insufficient data coverage and trends in available data generally not significant. An increase in 15 drought duration was observed with high confidence but medium confidence on the increase of drought 16 intensity (WGI AR6 Table 11.14, Seneviratne et al., 2021). Table 12.3 shows the estimates of changes in 17 land area per subregion affected by drought events, being this subregion which presented the highest changes 18 in CSA. 19 20 21 Table 12.3: Change in the percentage of land area affected by extreme drought in 2010-19, with respect to 1950-59 22 using the Standardised Precipitation-Evapotranspiration Index (SPEI); extreme drought is defined as SPEI -1.6 23 (Federal Office of Meteorology and Climatology MeteoSwiss, 2021). Data derived from Romanello et al. (2021). Average change in the percentage of land area in drought in 2010-19 with respect to 1950-59 Subregion At least 1 month in At least 3 months in At least 6 months in drought drought drought Central America (CA) 38.8% 17.6% 6.1% Northwest South America (NWS) 51.8% 25.3% 7.0% Northern South America (NSA) 52.5% 18.3% 2.5% South America Monsoon (SAM) 48.0% 34.4% 12.2% Northeast South America (NES) 64.5% 38.4% 12.0% Southeast SouthAmerica (SES) 16.4% 6.7% 0.4% Southwest South America (SWS) 20.5% 13.9% 7.5% Southern South America (SSA) -23.5% -8.8% -- 24 25 26 The projected warming for the extreme annual maximum temperatures over NES is TXx: +2°C for the 1.5°C 27 scenario and about +2.5°C for 2°C scenario (Hoegh-Guldberg et al., 2018). An increased number of tropical 28 nights with minimum temperatures exceeding the 20°C threshold is projected (Orlowsky and Seneviratne, 29 2012). In general, extreme heat will increase and cold spells decrease with high confidence. A decrease in 30 total precipitation is projected with high confidence with an increase in heavy precipitation events and an 31 increase in dryness (medium confidence). Increase in drought severity due to the combination of increased 32 temperatures, less rainfall, and lower atmospheric humidity (5 to 15% relative humidity reduction) create 33 water deficits, projected for the entire region after 2041 (3­4 mm day-1 reduction), particularly over western 34 NES and over the semiarid region (Marengo and Bernasconi, 2015; Marengo et al., 2017). Fire will 35 significantly increase (high confidence) (Figure 12.6). 36 37 12.3.5.2 Exposure 38 Do Not Cite, Quote or Distribute 12-28 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 NES is home to about 60 million people (estimate from IBGE (2019)), with >70% living in urban areas (data 2 from IBGE (2010); Silva et al. (2017)), and high poverty levels (> 50%, data from IBGE (2003)). People are 3 exposed to intense drought and famine (high confidence), and about 94% of the region has moderate to high 4 susceptibility to desertification (Marengo and Bernasconi, 2015; Spinoni et al., 2015; Vieira et al., 2015; 5 Mariano et al., 2018; Tomasella et al., 2018; Marengo et al., 2020c). The most severe dry spell of 2012­ 6 2013 affected about 9 million people, which were exposed to water, food and energy scarcity (Marengo and 7 Bernasconi, 2015). 8 9 People, infrastructure and economic activities are exposed to sea level rise in the 3800 km of coastline 10 (medium confidence). The high concentration of cities on the coast is a concern (Martins et al., 2017), with 11 all state capital cities but one on the coast, totalling almost 12 million exposed people (estimate from IBGE 12 (2019)). The ports of São Luís, Recife and Salvador are important exporters of Brazilian commodities, and 13 the beaches in the subregion are an international touristic destination, producing considerable revenues 14 (Pegas et al., 2015; Ribeiro et al., 2017). 15 16 Natural systems in NES are also exposed to climate change. In terrestrial ecosystems, 913,000 km2 of NES' 17 dry forest Caatinga vegetation (Silva et al., 2017) is exposed to predicted increase in dryness. Despite what 18 has been previously suggested, the Caatinga has high biodiversity and endemism (Silva et al., 2017), which 19 is exposed to habitat reduction due to climate change and agriculture expansion (Silva et al., 2019b). Fifty- 20 two percent of the freshwater fish (203 species) are endemic (Lima et al., 2017) and are exposed to predicted 21 reduction in river flow due to climate change (Marengo et al., 2017; de Jong et al., 2018). The coastal waters 22 contain a separate marine ecoregion due to its uniqueness (Spalding et al., 2007). The region is responsible 23 for 99% of the Brazilian shrimp production, exposed to sea level rise and increases in ocean temperature and 24 acidification (Gasalla et al., 2017). Most coral reefs in the Southern Atlantic Ocean are along NES's coast 25 (Leão et al., 2016), increasing its conservation and touristic value. The 685 km2 of coral reefs along NES's 26 coast (likely underestimated - Moura et al. (2013); UNEP-WCMC et al. (2018)) are exposed to increased 27 sea temperatures. 28 29 12.3.5.3 Vulnerability 30 31 NES is the world's most densely populated semi-arid land and its population is highly vulnerable to droughts 32 (high confidence), which have well-documented impacts on water and food security, human health and well- 33 being in the region (e.g., Confalonieri et al. (2014a); Marengo et al. (2017); Bedran-Martins et al. (2018)) 34 (Figure 12.7). The region's relative low economic development and poor social and health indicators 35 increase vulnerability, especially of poor farmers and traditional communities (Confalonieri et al., 2014a; 36 Bech Gaivizzo et al., 2019). In state capital cities, about 45% of the population live in poverty (data from 37 IBGE (2003)), often in slums with already deficient water supply and sewage systems and poor access to 38 health and education. Climate change will increase pressures on water availability, threatening water, energy 39 and food security (Marengo et al., 2017). 40 41 Natural systems in NES are also vulnerable (Figure 12.7). The Caatinga vegetation is particularly sensitive to 42 variations in water availability and climate change (Seddon et al., 2016; Rito et al., 2017; Dantas et al., 43 2020). It has already lost about 50% of its original vegetation cover (Souza et al., 2020), with only about 2% 44 of the remaining vegetation within fully protected areas (CNUC and MMA, 2020). Caatinga's high 45 vulnerability to climate change is further increased by the extensive conversion of native vegetation (high 46 confidence) (Rito et al., 2017; Silva et al., 2019b; Silva et al., 2019c). 47 48 Studies with terrestrial animals show that habitat loss increases the vulnerability of species to climate change 49 (high confidence) (de Oliveira et al., 2012; Arnan et al., 2018; da Silva et al., 2018b). NES' coral reefs have 50 shown some resilience to bleaching, but vulnerability is intensified by the synergism between chronic heat 51 stress caused by increased sea surface temperature (Teixeira et al., 2019) and other well-documented 52 stressors, such as coastal runoff, urban development, marine tourism, overexploitation of reef organisms and 53 oil extraction (high confidence) (Figure 12.8; Leão et al., 2016). 54 55 12.3.5.4 Impacts 56 Do Not Cite, Quote or Distribute 12-29 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Impacts of intense drought have been reported in NES since 1780, with severe losses in agricultural 2 production, livestock death, increase in agricultural prices, and human death (Figure 12.9; Marengo et al., 3 2017; Martins et al., 2019; Government of Brazil, 2020; Marengo et al., 2020c; Silva et al., 2020) (). The 4 rural population already suffers from natural water scarcity in the countryside. In 2012, the drought was 5 responsible for reducing up to 99% of the corn production in Pernambuco state (Government of Brazil, 6 2020). A predicted increase in drought, coupled with inadequate soil management practices by small farmers 7 and agribusiness, increases the region's susceptibility to desertification (Spinoni et al., 2015; Vieira et al., 8 2015; Mariano et al., 2018; Tomasella et al., 2018; Marengo et al., 2020c). In NES, 70,000 km2 have reached 9 a point at which agriculture is no longer possible (Government of Brazil, 2020). Intense droughts has 10 triggered migration to urban centres in and outside NES (Confalonieri et al., 2014a; Government of Brazil, 11 2020). More than 10 million people have been impacted by the drought of 2012/14 in the region, which was 12 responsible for water shortage and contamination, increasing death by diarrhoea (Marengo and Bernasconi, 13 2015; Government of Brazil, 2020). 14 15 There is growing evidence on the impacts of climate change on human health in NES, mostly linked to food 16 and water insecurity caused by recurrent long droughts (e.g., gastroenteritis and hepatitis) (high confidence) 17 (Figure 12.9; Sena et al., 2014; de Souza Hacon et al., 2019; Marengo et al., 2019; Government of Brazil, 18 2020; Salvador et al., 2020) . From 2071 to 2099, thermal conditions in NES might improve for vectors of 19 dengue, chikungunya and Zika (de Souza Hacon et al., 2019). Additionally, a high risk of mortality 20 associated with climatic stress in the period 2071­2099 is expected in São Francisco river basin (de Oliveira 21 et al., 2019; de Souza Hacon et al., 2019). 22 23 Recent studies predict strong negative impact of climate change on NES' agriculture (high confidence) 24 (Ferreira Filho and Moraes, 2015; Nabout et al., 2016; Gateau-Rey et al., 2018) (Figure 12.9; Table 12.4). 25 NES concentrates the bulk of the predicted loss of regional gross domestic product associated with 26 agriculture in Brazil (Ferreira Filho and Moraes, 2015; Forcella et al., 2015). Although agriculture gives a 27 modest contribution to the regions' economy, its drop could have a severe impact on the poorest rural 28 household, by shrinking the agricultural labour market and increasing food prices (Ferreira Filho and 29 Moraes, 2015; Government of Brazil, 2020). Expected increase in dryness is also predicted to impact the 30 region's hydroelectric power generation (Marengo et al., 2017; de Jong et al., 2018). Sea level rise has also 31 been reported to impact coastal cities such as Salvador, destroying urban constructions (Government of 32 Brazil, 2020). Sea level rise, increased ocean temperature and acidification may also negatively impact 33 NES's shrimp aquaculture production (Figure 12.8; Gasalla et al., 2017) . Along with climate change, 34 overfishing has driven exploited marine fish species to collapse (Verba et al., 2020). 35 36 Biodiversity in NES is highly threatened by climate change in terrestrial (medium confidence: medium 37 evidence, high agreement) and freshwater (low confidence: low evidence, high agreement) ecosystems 38 (Figure 12.9). There are few studies projecting the likely impact of climate change on NES' biodiversity, 39 especially on its endemic freshwater fish. Recent studies have already reported the reduction in several 40 endemic plant species affecting pollination and seed dispersal (Bech Gaivizzo et al., 2019; Cavalcante and 41 Duarte, 2019; Silva et al., 2019b). Studies with terrestrial animals predict that most groups would be 42 negatively impacted by climate change (de Oliveira et al., 2012; Arnan et al., 2018; da Silva et al., 2018b; 43 Montero et al., 2018). Changes in the abundance of coral reef community and extreme reduction in coral 44 cover have been observed in NES (de Moraes et al., 2019; Duarte et al., 2020). A number of observed coral 45 bleaching events associated with abnormal increase in sea temperatures have occurred in NES (Krug et al., 46 2013; Leão et al., 2016; de Oliveira Soares et al., 2019) (Figure 12.8), but thus far mortality remained low 47 and corals have been able return to normal values or remain stable after sea water temperature rise (medium 48 confidence: medium evidence, high agreement) (Leão et al., 2016). Mangroves in the region have shown 49 increased mortality, but have also expanded their range inland (Figure 12.6; Godoy and Lacerda, 2015; 50 Cohen et al., 2018) . Future projections include mangrove landward expansion and lower migration rates by 51 2100 (Cohen et al., 2018). 52 53 12.3.6 Southeast South America (SES) Sub-region 54 55 12.3.6.1 Hazards 56 Do Not Cite, Quote or Distribute 12-30 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 The increase in the intensity and frequency of hot extremes and decrease in the intensity and frequency of 2 cold extremes was observed with high confidence (Rusticucci et al., 2017; Wu and Polvani, 2017) (WGI 3 AR6 Table 11.13) (Seneviratne et al., 2021). There is low confidence that the decrease in hot extremes over 4 SES is related with an increase of extreme precipitation (Wu and Polvani, 2017). 5 6 Over SES most of the stations have registered an increase in annual rainfall, largely attributable to changes in 7 the warm season; this is one of few sub-regions where a robust positive trend in precipitation and significant 8 intensification of heavy precipitation has been detected since the beginning of the 20th century (high 9 confidence) but with medium confidence in a reduction of hydrological droughts (Vera and Díaz, 2015; 10 Saurral et al., 2017; Lovino et al., 2018; Avila-Diaz et al., 2020; Carvalho, 2020; Dereczynski et al., 2020; 11 Dunn et al., 2020; Marengo et al., 2020a; Olmo et al., 2020) (WGI AR6 Table 11.14) (Seneviratne et al., 12 2021). A higher observed frequency of extratropical cyclones in the region has been detected (Parise et al., 13 2009; Reboita et al., 2018) with three cyclogenetic foci: South-southeast Brazil, extreme south of Brazil and 14 Uruguay, and southeast of Argentina. 15 16 In Montevideo, mean sea-levels increased over the past 20 years, reaching 11 cm from 1902 to 2016, and a 17 recent accelerating trend has been observed (Gutiérrez et al., 2016b). A value of water-level rise and its 18 acceleration for Buenos Aires was calculated from a record of annual mean water levels obtained from 19 hourly levels (1905­2003). Annual mean water level showed a trend of +1.7 ± 0.05 mm yr-1, and an 20 acceleration of +0.019 ± 0.005 mm yr-2 (D'Onofrio et al., 2008). 21 22 Increasing trends in mean air temperature and extreme heat, and decreasing cold spells are projected (high 23 confidence) (WGI AR6 Table 12.6Ranasinghe et al., 2021). The increase in the frequency of warm nights is 24 larger than that projected for warm days consistent with observed past changes that have been related with 25 changes in cloud cover that affect differently daytime temperatures as compared to night time temperatures 26 (López-Franca et al., 2016; Menéndez et al., 2016; Feron et al., 2019). 27 28 Increases in mean precipitation (high confidence), pluvial floods and river floods are projected (medium 29 confidence) (Nunes et al., 2018) (WGI AR6 Table 12.6) (Ranasinghe et al., 2021). Droughts in the La Plata 30 Basin will be more frequent in the medium-term (2011-2040) and the distant future (2071-2100) (with 31 respect to the 1979-2008 period), but also shorter and more severe, for the more extreme emission scenario 32 (RCP8.5) (low confidence) (Carril et al., 2016). 33 34 Negative trend in the annual number of cyclone events in the long-term future of 3.6 to 6.5% (2070-2098) 35 are projected, that showed an increase of 3 to 11% (2080-2100 for the A1B scenario) (Grieger et al., 2014; 36 Reboita et al., 2018). All coastal and oceanic climate impact drivers (relative sea level, coastal flood and 37 erosion, marine heatwaves and ocean aridity) are expected to increase by mid-century in the RCP8.5 38 scenario (high confidence) (WGI AR6 Table 12.6, Ranasinghe et al., 2021). 39 40 12.3.6.2 Exposure 41 42 Higher temperatures and rising sea levels, changes in rainfall patterns, increased frequency and intensity of 43 extreme weather events, could generate risks to the energy and the infrastructure sectors, and to the mining 44 and metals network. In the Plata basin, urban floods have become more frequent, causing infrastructure 45 damage and sometimes substantial mortality (high confidence) (Barros et al., 2015; Zambrano et al., 2017; 46 Nagy et al., 2019; Mettler-Grove, 2020; Morales-Yokobori, 2021; Oyedotun and Ally, 2021). A large 47 increase in landslides and flash floods is also predicted for the Brazilian portion of SES, where they are 48 responsible for the majority of the deaths related to natural disasters in the country (high confidence) 49 (Debortoli et al., 2017; Haque et al., 2019; Saito et al., 2019; Marengo et al., 2020d; da Fonseca Aguiar and 50 Cataldi, 2021). Due to uncontrolled urban growth, 21.5 million people living in the large Brazilian cities of 51 São Paulo, Rio de Janeiro and Belo Horizonte (estimate from IBGE (2019)) are expected to be exposed to 52 water scarcity, despite great water availability in the region (medium evidence, medium agreement) 53 (Marengo et al., 2017; Lima and Magaña Rueda, 2018; Marengo et al., 2020b). 54 55 The expected increase in temperature also exposes the population in large cities to extreme heat. Urban heat 56 islands are already a reality in large cities in the region, such as Buenos Aires (high confidence) (Wong et al., 57 2013; Sarricolea and Meseguer-Ruiz, 2019; Wu et al., 2019; Mettler-Grove, 2020), Rio de Janeiro (high Do Not Cite, Quote or Distribute 12-31 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 confidence) (Ceccherini et al., 2016; Neiva et al., 2017; Geirinhas et al., 2018; Peres et al., 2018; Sarricolea 2 and Meseguer-Ruiz, 2019; Wu et al., 2019; de Farias et al., 2021) and São Paulo (high confidence) (Mishra 3 et al., 2015; Barros and Lombardo, 2016; Ceccherini et al., 2016; Vemado and Pereira Filho, 2016; de 4 Azevedo et al., 2018; Lima and Magaña Rueda, 2018; Ferreira and Duarte, 2019; Lapola et al., 2019a; 5 Sarricolea and Meseguer-Ruiz, 2019; Wu et al., 2019), with reported impact on human health in the latter 6 (medium confidence: medium evidence, medium agreement) (e.g., Araujo et al. (2015); Son et al. (2016); 7 Diniz et al. (2020)). These cities alone represent 22 million people exposed to increased heat (estimate from 8 IBGE (2019) and from INDEC (2010)). 9 10 The sub-region presents a high frequency of occurrence of intense severe convection events (Section 11 12.3.6.1). Because of this situation, strong winds from the south or southeast and high water levels affect the 12 whole Argentine coast, as well as the Rio de la Plata shores, Uruguay, and southern Brazil (Isla and Schnack, 13 2009). The coast of the Plata River is subject to flooding when there are strong winds from the southeast 14 (sudestadas). As sea level rises as a result of global climate change, storm surge floods will become more 15 frequent in this densely populated area, particularly in low-lying areas (high confidence) (Figure 12.8; 16 D'Onofrio et al., 2008; Nagy et al., 2014a; Santamaria-Aguilar et al., 2017; Nagy et al., 2019 impacts and 17 adaptation in Central and South America coastal areas; Cerón et al., 2021) . 18 19 The region's natural systems are also exposed to climate change. SES region houses two important 20 biodiversity hotspots, with high levels of species endemism: the Cerrado and the Atlantic Forest, where 21 about 72% of Brazil's threatened species can be found (PBMC, 2014). 22 23 12.3.6.3 Vulnerability 24 25 The Rio de la Plata basing and the city of Buenos Aires are highly vulnerable to recurring floods, and the 26 increasing number of newcomers to the area reduce the collective cultural adaptation developed by older 27 neighbours (high confidence) (Barros, 2006; Nagy et al., 2019; Mettler-Grove, 2020; Morales-Yokobori, 28 2021; Oyedotun and Ally, 2021). Extreme events, including storm surges and coastal inundation/flooding 29 caused injuries and economic/environmental losses on the urbanized coastline of Southern Brazil (States of 30 Sao Paulo and Santa Catarina) (high confidence) (Muehe, 2010; Khalid et al., 2020; Ohz et al., 2020; de 31 Souza and Ramos da Silva, 2021; Quadrado et al., 2021; Silva de Souza et al., 2021). 32 33 Cities like Rio de Janeiro and São Paulo are overpopulated, where most people live in poor conditions of 34 inadequate housing and sanitation, such as slums, with little and no trees and high temperatures. These 35 people have low access to sanitation, public health and residential cooling and are vulnerable to the effects of 36 heat islands on human comfort and health (Figure 12.7). These include cardiopulmonary and vector-borne 37 diseases, and even death (medium confidence: medium evidence, medium agreement) (Araujo et al., 2015; 38 Mishra et al., 2015; Geirinhas et al., 2018; Peres et al., 2018). Heat stress is known to worsen cardiovascular, 39 diabetic and respiratory conditions (Lapola et al., 2019a). As an effect of Heat islands, these people are also 40 vulnerable to injuries and casualties due to increased thunderstorms, causing economic losses and other 41 social problems (Vemado and Pereira Filho, 2016). 42 43 12.3.6.4 Impacts 44 45 Despite the observed increase in rainfall amount in the region, between 2014 and 2016 Brazil endured a 46 water crisis that affected the population and economy of major capital cities in the SES region Brazil 47 (Blunden and Arndt, 2014; Nobre et al., 2016a). Extremely long dry spells have become more frequent in 48 southeast of Brazil, affecting 40 million people and the economies in cities such as Rio de Janeiro, São Paulo 49 and Belo Horizonte, which are the industrial pole of the country (medium confidence: medium evidence, 50 medium agreement) (PBMC, 2014; Nobre et al., 2016a; Cunningham et al., 2017; Marengo et al., 2017; 51 Lima and Magaña Rueda, 2018; Marengo et al., 2020b). It also impacted agriculture, affecting food supply 52 and rural livelihoods, especially in Minas Gerais (Nehren et al., 2019). Agricultural prices increased by 30% 53 in some cases and harvest yields of sugar cane, coffee and fruits suffered a reduction of 15­40% in the 54 region. The number of fires increased by 150%, and energy prices increased by 20­25%, as most electricity 55 from hydroelectric power (Nobre et al., 2016a). In Argentina, projected changes in hydrology of Andean 56 rivers associated to glacier retreat are predicted to have negative impacts on the region's fruit production 57 (low evidence, medium agreement) (Barros et al., 2015). Do Not Cite, Quote or Distribute 12-32 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Heat islands affect ecosystems by increasing the energy consumption for cooling, the concentration of 3 pollutants and the incidence of fires (high confidence) (Wong et al., 2013; Akbari and Kolokotsa, 2016; 4 Singh et al., 2020b; Ulpiani, 2021). It also affects human health, as well increasing the incidence of 5 respiratory, cardiovascular diseases (medium confidence: medium evidence, medium agreement) (Araujo et 6 al., 2015; Barros and Lombardo, 2016; de Azevedo et al., 2018; Geirinhas et al., 2018). 7 8 Warming temperatures have been implicated in the emergence of dengue in temperate latitudes increasing 9 populations of Aedes aegypti (high confidence) (Natiello et al., 2008; Robert et al., 2019; Estallo et al., 2020; 10 Robert et al., 2020; López et al., 2021) (Table 12.1), and field studies have shown the role of local climate in 11 vector activity (Benitez et al., 2021). Figure 12.5 shows the modelled transmission suitability for dengue for 12 two climate change scenarios. Future increase in the number of months suited for transmission of dengue is 13 highest in SES (see SM12.8 for additional information). There is additional evidence of the spread of 14 arbovirus transmission into southern temperate latitudes (Basso et al., 2017), however a longer historical 15 time series is needed to understand climate-disease interactions, given the relatively recent emergence in this 16 region. 17 18 19 20 Figure 12.5: Predicted thermal suitability for transmission of dengue by Aedes aegypti mosquitoes, mapped as the 21 number of months of the year suitable under baseline or current conditions (2015), and in 2030, 2050, 2080 under two 22 representative concentration pathways, RCP4.5 and RCP8.5. Adapted from Ryan et al. (2019). See SM12.8 for 23 additional data on population at risk for dengue and Zika in the subregions and methodological details. 24 25 26 Sea-level rise impacted the port complex in Santa Catarina, which during the last six years interrupted its 27 activities 76 times due to strong winds or big waves with estimated losses varying between USD 25,000 and 28 USD 50,000 for each 24 idle hours (Ohz et al., 2020). Historically, extratropical cyclones associated with 29 frontal systems cause storm surges in Santos city. Although there are no fatality records, these events cause 30 several socio-economic losses, especially in vulnerable regions including the Port of Santos, the largest port Do Not Cite, Quote or Distribute 12-33 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 in Latin America (São Paulo). According to 88-year time span (1928-2016), the frequency of storm surge 2 events were three times more frequent in the last 17 years (2000-2016), than in the previous period of 71 3 years (1928-1999) (Souza et al., 2019). 4 5 There are many projected impacts of climate change on natural systems. The impacts of sea-level rise are 6 habitat destruction and the invasion of exotic species, affecting biodiversity and the provision of ecosystem 7 services (Figure 12.8; Nagy et al., 2019) (). 8 9 SES is a global priority for terrestrial biodiversity conservation, housing two important biodiversity 10 hotspots--the Atlantic Forest and Cerrado--which are among the World's most studied biodiversity-rich 11 spots in terms of climate change impact on biodiversity, especially for terrestrial vertebrates (Cross-Chapter 12 Paper 1.2.2; Manes et al., 2021) . An increasing number of studies show that the Atlantic Forest and Cerrado 13 are at risk of biodiversity loss, largely due to projected reduction of species' geographic distributions in 14 many different taxa (e.g., Loyola et al. (2012); Ferro et al. (2014); Loyola et al. (2014); Hoffmann et al. 15 (2015); Martins et al. (2015); Aguiar et al. (2016b); Vale et al. (2018); Borges et al. (2019); Braz et al. 16 (2019); Vale et al. (2021)). Cerrado savannas are projected to be the hotspot most negatively impacted by 17 climate change within South America, mostly though range contraction of plant species (very high 18 confidence), while the Atlantic Forest is projected to be highly impacted especially though the contraction of 19 the distribution of endemic species (very likely) (Cross-Chapter Paper 1.2.2; Figure 12.10; Manes et al., 20 2021) . Reductions in species' distribution are also projected in the La Plata Basin for subtropical 21 amphibians (Schivo et al., 2019) and the river tiger (Salminus brasiliensis), a keystone fish of economic 22 value (Ruaro et al., 2019). Farming of mussels and oysters in the region is predicted to be negatively 23 impacted by climate change, particularly sea-level rise, and ocean warming and acidification (Gasalla et al., 24 2017). Some more localized habitats are also at risk of losing area due to climate change, such as the 25 meadows of northwest Patagonia (Crego et al., 2014) and mangroves of southern Brazil (Godoy and 26 Lacerda, 2015). Predicted changes in global climate along with agricultural expansion will strongly affect 27 South American wetlands, which comprise around 20% of the continent and bring many benefits, such as 28 biodiversity conservation and water availability (Junk, 2013). 29 30 12.3.7 Southwest South America (SWS) Sub-region 31 32 12.3.7.1 Hazards 33 34 Significant increases in the intensity and frequency of hot extremes and significant decreases in the intensity 35 and frequency of cold extremes have likely been observed for the region (Skansi et al., 2013; Ceccherini et 36 al., 2016; Meseguer-Ruiz et al., 2018; Vicente-Serrano et al., 2018; Dereczynski et al., 2020; Dunn et al., 37 2020; Olmo et al., 2020) (WGI AR6 Table 11.13) (Seneviratne et al., 2021). In particular, a significant 38 increment in the duration and frequency of heatwaves mainly in central Chile from 1961 to 2016 has been 39 observed (Piticar, 2018). 40 41 A robust drying trend for Chile (30ºS­48°S) has been recorded (medium confidence) (Saurral et al., 2017; 42 Boisier et al., 2018). However, inconsistent trends over the region in the magnitude of precipitation extremes 43 with both decreases and increases (Chou et al., 2014; Giorgi et al., 2014; Heidinger et al., 2018; Meseguer- 44 Ruiz et al., 2018) (WGI AR6 Table 11.14) (Seneviratne et al., 2021) have been observed (low confidence). 45 The glacier equilibrium line altitude has presented an overall increase over central Chilean Andes (Barria et 46 al., 2019). 47 48 For central Chile, a significant increase (5% to 20% in the last 60 years) in wave heights in the sea has been 49 observed (Martínez et al., 2018). From 1982 to 2016, sea level at central Chile have increased 5 mm yr-1, 50 where El Niño events of 1982-1983 and 1997-1998 caused an extreme increase of 15 to 20 cm in the mean 51 sea level (Campos-Caba, 2016; Martínez et al., 2018). 52 53 From 1946 to 2017, the number of fires and areas burned have increased significantly in Chile (high 54 confidence) (González et al., 2011; Jolly et al., 2015; Úbeda and Sarricolea, 2016; de la Barrera et al., 2018; 55 Urrutia-Jalabert et al., 2018). Fires are attributed to changes in the temperatures regimes (González et al., 56 2011; de la Barrera et al., 2018; Gómez-González et al., 2018) and precipitation regimes (medium 57 confidence) (Gómez-González et al., 2018; Urrutia-Jalabert et al., 2018). Do Not Cite, Quote or Distribute 12-34 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 The glaciers of the Southern Andes (including the SWS and SSA regions) show the highest glacier mass loss 3 rates worldwide (high confidence) contributing to sea level rise (Jacob et al., 2012; Gardner et al., 2013; 4 Dussaillant et al., 2018; Braun et al., 2019; Zemp et al., 2019). Since 1985, the glacier area loss in the sub- 5 region is in a range of 20 up to 60% (Braun et al., 2019; Reinthaler et al., 2019b). 6 7 Four sets of downscaling simulations based on the Eta Regional Climate Model forced by two global climate 8 models (Chou et al., 2014) projected warmer conditions (more than 1°C) for all the sub-region by 2050 9 under the RCP4.5 scenario (medium confidence). Extremely warm December-January-February days as well 10 as the number of heatwaves per season are expected to increase by 5­10 times in the northern Chile (Feron et 11 al., 2019), likely increasing in the intensity and frequency of hot extremes over all the region (WGI AR6 12 Table 11.13) (Seneviratne et al., 2021). Drier conditions (medium confidence), by mean of the decrease of 13 total annual and extreme precipitations, are expected to increase for Southern Chile but inconsistent changes 14 in the sub-region (low confidence) (Chou et al., 2014) (WGI AR6 Table 11.14) (Seneviratne et al., 2021) 15 with high confidence on increase of fire weather and decrease of permafrost and snow extent (WGI AR6 16 Table 12.6, Ranasinghe et al., 2021). 17 18 Regional sea-level change for the region predicted by 2100 show that total mean SLR along the coast will lie 19 between 34 cm and 52 cm for the RCP4.5 scenario, and between 46 cm and 74 cm for the RCP8.5 scenario 20 with high confidence (Albrecht and Shaffer, 2016; WGI AR6 Table 12.6, Ranasinghe et al., 2021). 21 22 12.3.7.2 Exposure 23 24 There is high confidence that age and socio-economic status are key factors determining health exposure and 25 quality of life in SWS where low-income areas show an insufficient number of public spaces to provide 26 acceptable environmental quality in comparison with the high-income areas (Romero-Lankao et al., 2013; 27 Fernández and Wu, 2016; Paz et al., 2016; Hystad et al., 2019; Smith and Henríquez, 2019; Jaime et al., 28 2020; Pino-Cortés et al., 2020). 29 30 Profound social inequalities, urban expansion and the inadequate city planning (e.g., drainage network) 31 increase exposure to flooding events and landslides (high confidence) (Müller and Höfer, 2014; Rojas et al., 32 2017; Lara et al., 2018), heat hazards such as heatwaves (high confidence) (Welz et al., 2014; Qin et al., 33 2015; Inostroza et al., 2016; Welz and Krellenberg, 2016; Krellenberg and Welz, 2017), and the loss and 34 fragmentation of green infrastructure (Hernández-Moreno and Reyes-Paecke, 2018). SWS cities show the 35 highest levels of air pollution of CSA (medium confidence: medium evidence, high agreement) (Pino et al., 36 2015; Huneeus et al., 2020; González-Rojas et al., 2021), where the state air quality alerts have limited effect 37 on protective health behaviours, being the public perception about air pollution highly dissimilar among the 38 population (Boso et al., 2019). In particular, human communities living in coastal cities show a negative 39 safety perception about the performance of the infrastructure and coastal defences to flood events (low 40 confidence) (González and Holtmann-Ahumada, 2017; Igualt et al., 2019). 41 42 Although climate change is critically important for the current and future status of mining activity in SWS 43 (Odell et al., 2018), and SWS areas subjected to mining activities are highly exposed to water risk (Northey 44 et al., 2017), to date, there is low evidence of climate change impacting mining activities (Corzo and 45 Gamboa, 2018; Odell et al., 2018). 46 47 12.3.7.3 Vulnerability 48 49 Rapid changes in temperature and precipitation regimes make terrestrial ecosystems highly vulnerable to 50 climate change (high confidence) (Salas et al., 2016; Fuentes-Castillo et al., 2020) (Figure 12.7). Terrestrial 51 ecosystems dominated by exotic species (e.g., pine) with lower landscape heterogeneity, degraded soils and 52 close to settlements and roads are highly vulnerable to wildfires in comparison to forests dominated by 53 native trees (high confidence) (Altamirano et al., 2013; Castillo-Soto et al., 2013; Cóbar-Carranza et al., 54 2014; Salas et al., 2016; Bañales-Seguel et al., 2018; Gómez-González et al., 2018; Sarricolea et al., 2020). 55 Changes in the land use, artificial forestation, deforestation, agricultural abandonment and urbanization have 56 provoked a permanent degradation of old-growth forest putting at risk the biodiversity, recreation and 57 ecotourism (medium confidence: medium evidence, high agreement) (Rojas et al., 2013; Nahuelhual et al., Do Not Cite, Quote or Distribute 12-35 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2014). Marine coastal ecosystems such as dunes, sandy beaches and wetlands show a high deterioration 2 decreasing the ability to mitigate extreme events (medium confidence: low evidence, high agreement) 3 (González and Holtmann-Ahumada, 2017; Ministerio de Medio Ambiente de Chile, 2019). 4 5 Water sector shows a very high vulnerability (high confidence) (Figure 12.7) mainly due the weak water 6 governance focused on market aspects (e.g., inter-sectoral water transactions, setting rates, granting 7 concessions, waiving the water right) (high confidence) (Hurlbert and Diaz, 2013; Valdés-Pineda et al., 8 2014; Barría et al., 2019; Hurlbert and Gupta, 2019; Muñoz et al., 2020a; Urquiza and Billi, 2020b). Potable 9 water and adequate sanitization is available in SWS; however, water availability along Chile is unevenly 10 distributed in rural communities (high confidence) (Valdés-Pineda et al., 2014; Nelson-Nuñez et al., 2019). 11 Spatial differences on water availability are enhanced by the strong population growth, economic 12 development, mining activities, and the high dependence of agriculture to irrigation (high confidence) 13 (Stathatou et al., 2016; Northey et al., 2017; Fercovic et al., 2019). Droughts in SWS are a major threat to 14 water security (high confidence) (Aitken et al., 2016; Núñez et al., 2017) as river streamflow are highly 15 dependent on the inter-annual to decadal climate conditions, snow melting processes, rainfall events (Boisier 16 et al., 2016), and impacted by land uses and changes in irrigated agriculture (medium confidence: medium 17 evidence, high agreement) (Vicuña et al., 2013; Fuentes et al., 2021). 18 19 Energy and water needs of large-scale mining activities make this socio economic sector particularly 20 vulnerable to climate change; additionally, the relative lack of power of resource-poor communities living in 21 areas where such mining is making claims on water and energy resources renders these communities even 22 more vulnerable (Odell et al., 2018). Given new conditions generated by changes in a growing demand and 23 climate change, mining industries will need to increase resilience to extreme events; additionally, the 24 declining concentrations of mineral of interest in the raw material require greater energy input for extraction 25 and processing and new methods to avoid associated emissions are required (Hodgkinson and Smith, 2018). 26 27 Urban and agriculture sectors are vulnerable to climate change (medium confidence: medium evidence, high 28 agreement) (Figure 12.7) increasing problems and demand for water (high confidence) (Monsalves-Gavilán 29 et al., 2013; Meza et al., 2014; Fercovic et al., 2019). Important health problems (e.g., pathogenic infections, 30 changes in vector-borne diseases, mortality by heat, lower neurobehavioral performance, among others) have 31 been associated with agriculture, mining and thermal power production activities along SWS (high 32 confidence) (Muñoz-Zanzi et al., 2014; Valdés-Pineda et al., 2014; Pino et al., 2015; Cortés, 2016; 33 Berasaluce et al., 2019; Muñoz et al., 2019a; Ramírez-Santana et al., 2020). 34 35 The large-scale agricultural growth has increased the vulnerability to climate change by favouring the 36 detriment of traditional agriculture, the homogenization of the biophysical landscape and the replacement of 37 traditional crops and native forests with exotic species like pines and eucalyptus (high confidence) (Torres et 38 al., 2015) where farmers' climate change perception is highly dependent on the education level and the 39 access to meteorological information (low confidence) (Roco et al., 2015). Agricultural systems owned by 40 Indigenous Peoples (i.e., Mapuche, Quechua and Aymara farmers) seem to present lower vulnerability to 41 drought and higher response capacity than non-indigenous farmers thanks to the use of the traditional 42 knowledge of specific management techniques and the tendency to conserve species or varieties of crops 43 tolerant to water scarcity (low confidence) (Montalba et al., 2015; Saylor et al., 2017; Meldrum et al., 2018). 44 Fishery and aquaculture-related livelihoods are vulnerable to climate and non-climate drivers (medium 45 confidence: medium evidence, high agreement) such as sea surface warming and precipitation reduction 46 (Handisyde et al., 2017; Soto et al., 2019; González et al., 2021), changes in upwelling intensity (low 47 confidence) (Oyarzún and Brierley, 2019; Ramajo et al., 2020), eutrophication and harmful algal bloom 48 (HAB) events (Almanza et al., 2019), the lack of observational elements and data management (Garçon et 49 al., 2019), and events such as earthquakes and tsunamis (Marín, 2019). 50 51 Chile has experienced an accelerated economic growth which has reduced poverty, however important 52 geographical, economic and educational inequalities are still present (Repetto, 2016). Chilean healthcare 53 system has become more equitable and responsive to the population necessities (e.g., Health reform AUGE 54 program); however, the high relative inequalities in terms of income (OECD, 2018), education level, and the 55 rural­urban factor are determinants of the quality of care, the health system barriers, and the health 56 differential access (high confidence) (Frenz et al., 2014). Exposure and vulnerability to psychosocial risks in 57 SWS shows significant inequalities to natural disasters such as earthquakes according to socio-economic, Do Not Cite, Quote or Distribute 12-36 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 geographic and gender factors (high confidence) (Labra, 2002; Vitriol et al., 2014; Quijada et al., 2018) 2 which are increased by the absence of local planning and drills and the lack of coordination (Vitriol et al., 3 2014). Indigenous Peoples have the highest levels of vulnerability in Chile in terms of income, basic needs, 4 and access to services to climate change (low confidence) (Parraguez-Vergara et al., 2016). 5 6 12.3.7.4 Impacts 7 8 Increasing temperatures in SWS have impacted temperate forests (high confidence) (Peña et al., 2014; 9 Urrutia-Jalabert et al., 2015; Camarero and Fajardo, 2017; Fontúrbel et al., 2018; Venegas-González et al., 10 2018b; Peña-Guerrero et al., 2020). Increasing temperatures and decreasing precipitations have increased the 11 impacts of wildfires on terrestrial ecosystems (high confidence) (Boisier et al., 2016; Díaz-Hormazábal and 12 González, 2016; Martinez-Harms et al., 2017; de la Barrera et al., 2018; Gómez-González et al., 2018; 13 Urrutia et al., 2018; Bowman et al., 2019), creating conditions for future landslides and floods (de la Barrera 14 et al., 2018). 15 16 Future projections show important changes in the productivity, structure and biogeochemical cycles in SWS 17 temperate and rainforests (medium confidence: medium evidence, high agreement) (Gutiérrez et al., 2014; 18 Correa-Araneda et al., 2020), and their fauna (low confidence) (Glade et al., 2016; Bourke et al., 2018). The 19 "Chilean Winter Rainfall-Valdivian Forests'' is a biodiversity-rich spot (Manes et al., 2021) (Cross-Chapter 20 Paper 1.2.2) projected to suffer habitat change, with loss of vegetation cover in the future due to climate 21 change (medium confidence: medium evidence, high agreement) (Jantz et al., 2015; Mantyka-Pringle et al., 22 2015). Species are projected to suffer changes in their distribution, including decrease in climatic refugia for 23 vertebrates (low confidence) (Cuyckens et al., 2015; Warren et al., 2018). 24 25 Increasing temperatures have enlarged the number and area extent of glacier lakes in Central Andes, 26 Northern Patagonia and Southern Patagonia (high confidence) (Wilson et al., 2018), while decreased rainfall 27 and rapid glacier melting have provoked changes in the environmental, biogeochemical and biological 28 properties of the central-southern and Andes Chilean lakes (low confidence) (Pizarro et al., 2016). 29 30 Increasing glacier lake outburst floods (GLOF), ice and rock avalanches, debris flows, and lahars from ice- 31 capped volcanoes have been observed in SWS (Iribarren Anacona et al., 2015; Jacquet et al., 2017; 32 Reinthaler et al., 2019b). There is low evidence about the effects of warming and degrading permafrost on 33 slope instability and landslides in these regions (Iribarren Anacona et al., 2015). 34 35 Increasing temperatures, decreasing precipitation regimes, and an unprecedented long-term drought have 36 decreased the annual average rivers streamflow that supply SWS megacities such as Santiago (high 37 confidence) (Meza et al., 2014; Muñoz et al., 2020a), with important and negative effects over the water 38 quality (Bocchiola et al., 2018; Yevenes et al., 2018) threatening irrigated agriculture activities (medium 39 confidence: medium evidence, high agreement) (Yevenes et al., 2018; Oertel et al., 2020; Peña-Guerrero et 40 al., 2020). Large reductions in the groundwater availability of the SWS region (Meza et al., 2014) and a 41 sustained decreasing of the mean annual flows (Ragettli et al., 2016; Bocchiola et al., 2018), especially 42 during the snowmelt season (Vargas et al., 2013) have been observed in SWS. Drought has affected wetlands 43 (low confidence) (Zhao et al., 2016; Domic et al., 2018), and desert ecosystems (medium confidence: medium 44 evidence, high agreement) (Acosta-Jamett et al., 2016; Neilson et al., 2017; Díaz et al., 2019). 45 46 There is low evidence about shoreline retreat attributed to climate change (Martínez et al., 2018; Ministerio 47 de Medio Ambiente de Chile, 2019) although increasing wind intensity along the central Chilean coast has 48 caused important damages in the coastal infrastructure and buildings (Winckler et al., 2017) and changes of 49 seawater properties and processes (low confidence) (Schneider et al., 2017; Aguirre et al., 2018). Ocean and 50 coastal ecosystems in SWS are sensitive to upwelling intensity which affect the abundance, diversity, 51 physiology and survivorship of coastal species (high confidence) (Anabalón et al., 2016; Jacob et al., 2018; 52 Ramajo et al., 2020) (Figure 12.8). Increasing radiation and temperatures, and reduced precipitations in 53 conjunction with increased nutrient load have increased HAB events producing massive fauna mortalities 54 (high confidence) (León-Muñoz et al., 2018; IPCC, 2019b, SPM A8.2 and B8.3; Quiñones et al., 2019; Soto 55 et al., 2019; Armijo et al., 2020). Multiple resources subjected to fisheries and aquaculture are highly 56 vulnerable to storms, alluvial disasters, ocean warming, ocean acidification, increasing ENSO extreme 57 events, and lower oxygen availability (high confidence) (Figure 12.8; García-Reyes et al., 2015; Silva et al., Do Not Cite, Quote or Distribute 12-37 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2015; Duarte et al., 2016; Lagos et al., 2016; Navarro et al., 2016; Lardies et al., 2017; Duarte et al., 2018; 2 IPCC, 2019b; Mellado et al., 2019; Ramajo et al., 2019; Silva et al., 2019a; Bertrand et al., 2020). Ocean and 3 coastal ecosystems, especially the EEZ will be highly impacted by climate change in the near and long-term 4 (high confidence) (Figure 12.8; Table SM12.3; Silva et al., 2015; Silva et al., 2019a). 5 6 Changes in the temperature and drought have impacted crops significantly (medium confidence: medium 7 evidence, high agreement) (Ray et al., 2015; Zambrano et al., 2016; Lesjak and Calderini, 2017; Ferrero et 8 al., 2018; Piticar, 2018; Haddad et al., 2019; Zúñiga et al., 2021). Table 12.4 shows the changes in crop 9 growth duration, which affect the yields. Higher negative numbers then indicate yield reduction for the crop. 10 Increasing temperatures and decreasing precipitation are expected to impact the agriculture sector (i.e., fruits 11 crops, and forests) across the entire sub-region with the largest impacts in the northern and central zone (high 12 confidence) (Mera et al., 2015; Zhang et al., 2015; Silva et al., 2016; Lizana et al., 2017; Reyer et al., 2017; 13 Toro-Mujica et al., 2017; Beyá-Marshall et al., 2018; Lobos et al., 2018; O'Leary et al., 2018; Aggarwal et 14 al., 2019; Ávila-Valdés et al., 2020; Fernandez et al., 2020; Melo and Foster, 2021). Observed impacts and 15 future projections warn that increasing temperatures and decreasing precipitation will largely impact on 16 water demand by agricultural sectors (high confidence) (Novoa et al., 2019; Peña-Guerrero et al., 2020; 17 Webb et al., 2020). Extreme climate events have provoked that Indigenous Peoples (e.g., Mapuche, Uru and 18 Aymara) suffer scarcity of water, reduction of agricultural production, and a displacement of their traditional 19 knowledge and practices (medium confidence: low evidence, high agreement) (Parraguez-Vergara et al., 20 2016; Meldrum et al., 2018; Perreault, 2020). 21 22 23 Table 12.4: Average percentage change in crop growth duration for the period 2015-19. Crop growth duration refers to 24 the time taken in a year for crops to accumulate the reference period (1981-2010) average growing season Accumulated 25 Temperature Total (ATT). As temperatures rise, the ATT is reached earlier (higher negative changes), the crop matures 26 too quickly, and thus yields are lower. "No data" means no data is available for the growth of that crop, in the specified 27 region. NP means that the crop is not present in significant areas in that region. Data derived from Romanello et al. 28 (2021). Region Winter wheat Spring Rice Maize Soybean wheat Central America (CA) -4.8% No data -1.9% -5.0% -4.7% Northwest South America (NWS) -3.8% -5.2% -5.2% -5.6% -3.1% Northern South America (NSA) NP NP -0.7% -3.1% 0.0% South America Monsoon (SAM) -5.3% -0.7% -1.4% -2.9% -1.5% Northeast South America (NES) -1.0% -1.3% -0.7% -3.5% -2.6% Southeast South America (SES) -2.3% -3.5% -2.3% -2.4% -2.7% Southwest South America (SWS) -2.3% -5.2% -10.0% -5.2% No data Southern South America (SSA) -0.8% -6.5% No data -1.6% No data 29 30 31 SWS cities have been largely impacted by wildfires, water scarcity and landslides affecting highways and 32 local roads, as well as, potable water supply (Sepúlveda et al., 2015; Araya-Muñoz et al., 2016). Increasing 33 temperature and heat extreme events in cities have increased the demand for water, the damage of urban 34 infrastructure (Monsalves-Gavilán et al., 2013), and accelerated the ageing and the death of trees (high 35 confidence) (Moser-Reischl et al., 2019). Increasing temperature will modify the energy demand in cities in 36 northern and central Chile (Rouault et al., 2019). 37 38 Increasing temperature, heat extreme events and air pollution in SWS have significantly impacted the 39 population health (cardiac complications, heat stroke, and respiratory diseases) (high confidence) (Table 40 12.2; Leiva G et al., 2013; Monsalves-Gavilán et al., 2013; Pino et al., 2015; Herrera et al., 2016; Henríquez 41 and Urrea, 2017; Ugarte-Avilés et al., 2017; de la Barrera et al., 2018; Johns et al., 2018; Bowman et al., Do Not Cite, Quote or Distribute 12-38 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2019; González et al., 2019; Matus C and Oyarzún G, 2019; Sánchez et al., 2019; Terrazas et al., 2019; 2 Cakmak et al., 2021; Zenteno et al., 2021).There is low confidence about area changes of Chagas disease 3 (Tapia-Garay et al., 2018; Garrido et al., 2019), and transmission rates in the future (Ayala et al., 2019). 4 5 12.3.8 Southern South America (SSA) Sub-region 6 7 12.3.8.1 Hazards 8 9 There were inconsistent trends and insufficient data coverage about extreme temperatures and precipitation 10 (low confidence) but with medium confidence an increase in the frequency of meteorological droughts was 11 observed (Dereczynski et al., 2020; Dunn et al., 2020; WGI AR6 Tables 11.13, 11.14, 11.15, Seneviratne et 12 al., 2021; WGI AR6 Table 12.3, Ranasinghe et al., 2021). An increase in precipitation in Trelew, no change 13 for Comodoro Rivadavia, both stations located at Eastern Patagonia, and negative trends in austral summer 14 rainfall in southern Andes were observed (Vera and Díaz, 2015; Saurral et al., 2017). Chile's wildfires in 15 Patagonia (fire frequency and intensity) have grown at an alarming rate (Úbeda and Sarricolea, 2016). 16 Decreasing rainfall pattern in Punta Arenas is closely associated with the variability at inter-annual to inter- 17 decadal time scales of the main forcing system for climate in Patagonia. Snow Cover Extension (SCE) and 18 Snow Cover Duration decreased by an average of ~13 ± 2% and 43 ± 20 days respectively from 2000 to 19 2016, due to warming rather than drying (Rasmussen et al., 2007). In particular, the analysis of spatial 20 pattern of SCE indicates a slightly greater reduction on the eastern side (~14 ± 2%) of the Andes Cordillera 21 compared to the western side (~12 ± 3%). The longest time series of glacier mass balance data in the 22 Southern Hemisphere, the Echaurren Norte Glacier, lost 65% of its original area in the period 1955­2015 23 and disaggregated into two ice bodies in the late 1990s (Malmros et al., 2018; Pérez et al., 2018). 24 25 Mean temperatures in the SSA sub-region are projected to continue to rise up to +2.5°C in 2080 with respect 26 to the present climatology (Kreps et al., 2012). A rise in temperature means that the isotherm of 0°C will 27 move up the mountains leaving less surface for accumulation of snow (Barros et al., 2015). 28 29 An increase in the intensity and frequency of hot extremes and a decrease in the intensity and frequency of 30 cold extremes is likely projected (WGI AR6 Table 11.13, Seneviratne et al., 2021); CMIP6 models project an 31 increase in the intensity and frequency of heavy precipitation (medium confidence). 32 33 It is expected that an increase in the intensity of heavy precipitation, droughts and fire weather will intensify 34 through the 21st century in SSA but mean wind will decrease (medium confidence) (Kitoh et al., 2011; WGI 35 AR6 Tables 11.14 and Table 11.15, Seneviratne et al., 2021). The probability of having extended droughts, 36 such as the recently experienced mega-drought (2010-2015), increases to up to 5 events/100 yr (Bozkurt et 37 al., 2017). Snow, glaciers, permafrost and ice sheets will decrease with high confidence (WGI AR6 Table 38 12.6, Ranasinghe et al., 2021). The observed area and the elevation changes indicate that the Echaurren 39 Norte Glacier may disappear in the coming years if negative mass balance rates prevail (medium confidence) 40 (Farías-Barahona et al., 2019). 41 42 12.3.8.2 Exposure 43 44 Grasslands make a significant contribution to food security in Patagonia through providing part of the feed 45 requirements of ruminants used for meat, wool and milk production. There is a lack of information regarding 46 the combined effect of climate change and overgrazing and the consequences for pastoral livelihoods that 47 depend on rangelands. Temperature and the amount and seasonal distribution of precipitation were important 48 controls of vegetation structure in Patagonian rangelands (Gaitán et al., 2014). They found that over two- 49 thirds of the total effect of precipitation on above-ground net primary production (ANPP) was direct, and the 50 other third was indirect (via the effects of precipitation on vegetation structure). Thus, if evapotranspiration 51 and drought stress increase as temperature increases and rainfall decrease in water-limited ecosystems, it 52 would be expected a greater exposure of ranchers due to a reduction of stocking rate and therefore families' 53 income (medium confidence). The number of farmers (mainly family enterprises) exposed to climatic 54 hazards (drought) is approximately 70­80 thousand that have 14­15 million sheep in Argentina (Peri et al., 55 2021). 56 Do Not Cite, Quote or Distribute 12-39 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Argentinian Patagonia main cities have developed as the result of oil and gas extraction, which demand 2 massive quantities of water due to fracking and drilling techniques. Vaca Muerta is the major region in South 3 America where those techniques are used to extract oil and gas, and this will lead to an exacerbation of 4 current water scarcity and to competition with irrigated agriculture (Rosa and D'Odorico, 2019) which in the 5 context of drought may exacerbate socio-environmental conflicts (medium confidence). 6 7 12.3.8.3 Vulnerability 8 9 There are reports related to a decrease in survival, growth and higher vulnerability to drought and fire- 10 severity for species of native forest due to climate change and wildfire (high confidence) (Mundo et al., 11 2010; Landesmann et al., 2015; Whitlock et al., 2015; Jump et al., 2017; Camarero et al., 2018; Venegas- 12 González et al., 2018a). There is a reported coincidence between major changes in regional decline in the 13 growth of forests with severe droughts due to climatic variations over northern Patagonia (Rodríguez-Catón 14 et al., 2016). Once the forest decline begins, other contributing factors such as insects (e.g., defoliator 15 outbreaks) increase the forest vulnerability or accelerate the loss of forest health of previously stressed trees 16 (Piper et al., 2015). This region hosts unique temperate rainforests and it is particularly rich in endemic and 17 long-lived conifer species (e.g., Fitzroya cupressoides), which may be vulnerable to declines in soil moisture 18 availability (Camarero and Fajardo, 2017). Patagonia will probably be vulnerable by a decrease in 19 precipitation regimes due to climate change, and consequently many species that rely on meadows in an arid 20 environment will also be impacted (Crego et al., 2014). The floods triggered by strong ENSOs caused 21 significant changes in the crop production (Isla et al., 2018). 22 23 The development of various human activities and water infrastructure are decreasing water sources, changing 24 river basins from exoreic to endoreic and the disappearance of one lake in 2016 (Scordo et al., 2017). 25 Numerous dams for irrigation, some also used for hydropower, have been and are planned to be built despite 26 wind power generation potential (Silva, 2016). Oil and gas have played an important role in the rise of 27 Neuquén-Cipolletti as Patagonia's most populous urban area, and in the growth of Comodoro Rivadavia, 28 Punta Arenas, and Rio Grande, as well. 29 30 12.3.8.4 Impacts 31 32 The potential impact of climate change is of special concern in arid and semi-arid Patagonia, a >700,000 km2 33 region of steppe-like plains in Argentina. Thus, melting snow and ice in the glaciers of Patagonia and the 34 Andes will alter surface runoff into interior wetlands; sea level rise of between 20 and 60 cm will destroy 35 coastal marshes; and an increase in extreme events, such as storms, floods, and droughts, will affect 36 biodiversity in wet grasslands (medium confidence: low evidence, high agreement) (after Junk et al. 2013; 37 Joyce et al. 2016). Three species of lizard from Patagonia are at risk of extinction as a result of global 38 warming (Kubisch et al., 2016). 39 40 Patagonian ice fields in South America are the largest bodies of ice outside of Antarctica in the southern 41 hemisphere. They are losing volume due partly to rapid changes in their outlet glaciers which end up in lakes 42 or the oceans, becoming the largest contributors to eustatic sea level rise (SLR) in the world, per unit area 43 (Foresta et al., 2018; Moragues et al., 2019; Zemp et al., 2019). Most calving glaciers in the Southern 44 Patagonia ice field retreated during the last century (high confidence). Upsala Glacier retreat generated slope 45 instability and a landslide movement destroyed the western edge in 2013. The Upsala Argentina Lake has 46 become potentially unstable and may generate new landslides (Moragues et al., 2019). The climate effect on 47 the summer stratification of piedmont lakes is another issue in relation to glacier dynamics (Isla et al., 2010). 48 49 Between 41º and 56° South latitude, the absolute glacier area loss was 5450 km2 (19%) in the last 150 50 years, with an annual area reduction increase of 0.25% a-1 for the period 2005­2016 (Meier et al., 2018). The 51 small glaciers in the north of the Northern Patagonian Ice field had over all periods the highest rates of 52 0.92% a-1. In this sub-region, increased melting of ice is leading to changes in the structure and functioning 53 of river ecosystems and in freshwater inputs to coastal marine ecosystems (medium confidence: low 54 evidence, high agreement) (Aguayo et al., 2019). In addition, in the case of coastal areas, the importance of 55 tides and rising sea levels in the behaviour of river floods has been demonstrated (Jalón-Rojas et al., 2018). 56 Do Not Cite, Quote or Distribute 12-40 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Suitable areas for meadows (very productive areas for livestock production) will decrease by 7.85% by 2050 2 given predicted changes in climate (low confidence) (Crego et al., 2014). 3 4 A major drought from 1998 to 1999 coincident with a very hot summer led to extensive dieback in a 5 Nothofagus species (Suarez et al., 2004). In another dominant Nothofagus species, several periodic droughts 6 have triggered forest decline as of the 1940s (Rodríguez-Catón et al., 2016). 7 8 Climate change impacted ocean ecosystems by reducing kelps coverage, increasing reproductive failure and 9 chick mortality of penguins, and poleward expansion of saltmarshes in the Atlantic Patagonia. SSA houses 10 the Patagonian Steppe Global-200 terrestrial ecoregion being a conservation priority at global scale, but with 11 a clear lack of studies on likely future climate change impacts (Cross-Chapter Paper 1.2.2.2; Manes et al., 12 2021). The Patagonian Steppe may suffer pronounced expansion in invasive species' ranges under climate 13 change (low confidence) (Wang et al., 2017a). 14 15 Fire has been found to promote or halt biological invasions (medium confidence: medium evidence, high 16 agreement). For example, an analysis of Pinus spreading after wildfires in Patagonia reveals that there is a 17 high risk of pines becoming invasive if ignition frequency increases as a result of climate change (Raffaele et 18 al., 2016). According to Inostroza et al. (2016), the Magellan Region is one of the most fragile regions in 19 Patagonia and despite its low population densities, it is under a silent process of anthropogenic alteration 20 where between 53.1% and 68.1% of the area needs to be considered as influenced by human activity whom 21 are occupying pristine ecosystems even extensive conservation designations (Inostroza et al., 2016). Fire 22 exposure can result in several health problems for human populations; Table 12.5 shows that SSA is the 23 region with the highest exposure to wildfire danger. 24 25 26 Table 12.5: Change in population-weighted exposure to very high or extremely high wildfire risk. Data derived from 27 the Fire Danger Indices FDI produced by the Copernicus Emergency Management Service for the European Forest Fire 28 Information System EFFIS (available at Copernicus Emergency Management Service (2021)). High and very high 29 wildfire danger defined as FDI >= 5. Data derived from Romanello et al. (2021). Population-weighted mean days of exposure to extremely high and very high wildfire danger Subregion In 2001-04 In 2017-20 Change from 2001-04 to 2017-20 Central America (CA) 30.4 26.9 -3.5 Northwest South America (NWS) 4.2 4.6 0.5 Northern South America (NSA) 19.7 21.2 1.5 South America Monsoon (SAM) 16.0 27.8 11.8 Northeast South America (NES) 47.9 53.3 5.4 Southeast SouthAmerica (SES) 4.2 8.2 4.0 Southwest SouthAmerica (SWS) 31.9 58.4 26.5 Southern South America (SSA) 88.7 104.9 16.2 30 31 Do Not Cite, Quote or Distribute 12-41 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Figure 12.6: Observed trends (WGI AR6 Tables 11.13, 11.14, 11.15) (Seneviratne et al., 2021) and summary of 3 confidence in direction of projected change in climatic impact-drivers, representing their aggregate characteristic 4 changes for mid-century for scenarios RCP4.5, SSP3-44 4.5, SRES A1B, or above within each AR6 region, 5 approximately corresponding (for CIDs that are independent of sea-level rise) to global warming levels between 2°C 6 and 2.4°C (WGI AR6 Table 12.6) (Ranasinghe et al., 2021). 7 8 Do Not Cite, Quote or Distribute 12-42 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Figure 12.7: Sectoral distribution of vulnerability levels to climate change for the subregions. The vulnerability levels 3 are based on studies that include: i) databases with climate change vulnerability indexes by country and sector, ii) 4 researches that implement climate change vulnerability indexes by sector at the local, national, regional or global scale, 5 and iii) studies that define some vulnerability level based on the authors' expert judgment. Panel (a) shows the 6 vulnerability and confidence levels for each subregion. Panel (b) indicates the references used and the level of 7 vulnerability attributed by subregion. The numbers within the table indicate the reference used for the assessment in the 8 following order: 1) Aitken et al. (2016); 2) Anderson et al. (2018b); 3) Bañales-Seguel et al. (2018); 4) Bouroncle et al. 9 (2017); 5) CAF (2014); 6) Carrão et al. (2016); 7) Donatti et al. (2019); 8) Eguiguren-Velepucha et al. (2016); 9) FAO 10 (2020a); 10) FAO (2020b); 11) FAO (2021a); 12) FAO (2021b); 13) FAO (2021c); 14) FAO et al. (2021); 15) FAO and Do Not Cite, Quote or Distribute 12-43 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 ECLAC (2020); 16) Ferreira Filho and Moraes (2015); 17) Filho et al. (2016); 18) Fuentes-Castillo et al. (2020); 19) 2 FSIN and Global Network Against Food Crisis (2021); 20) Global Health Security Index (2019); 21) Godber and Wall 3 (2014); 22) Handisyde et al. (2017); 23) Hannah et al. (2017); 24) Immerzeel et al. (2020); 25) Inform Risk Index 4 (2021); 26) Koutroulis et al. (2019); 27) Krishnamurthy et al. (2014); 28) Lapola et al. (2019a); 29) Li et al. (2018); 30) 5 Lin et al. (2020); 31) Mansur et al. (2016); 32) Martins et al. (2017); 33) Menezes et al. (2018); 34) Nagy et al. 6 (2018);35) ND-Gain (2020); 36) Northey et al. (2017); 37) Olivares et al. (2015); 38) Pacifici et al. (2015); 39) Qin et 7 al. (2020); 40) Romeo et al. (2020); 41) Liu and Chen (2021); 42) Silva et al. (2019b); 43) Soto Winckler and Del 8 Castillo Pantoja (2019); 44) Soto et al. (2019); 45) Tomby and Zhang (2019); 46) Venegas-González et al. (2018b); 47) 9 Yeni and Alpas (2017); 48) Marengo et al. (2017); 49) Bedran-Martins et al. (2018); 50) Confalonieri et al. (2014a). 10 Detailed methodology can be found in SM12.2. 11 12 13 14 Figure 12.8: Climate and non-climate sensitivity drivers of ocean, coastal ecosystems and Exclusive Economic Zones 15 (EEZs) of Central and South America. 16 17 18 12.4 Key Impacts and Risks 19 20 This section synthesizes key risks across the Central and South America CSA region. It follows the 21 definition and concept of risk provided in AR5, distinguishing the risk components, climatic hazard, 22 exposure and vulnerability of people and assets (IPCC, 2014). This concept is further developed in AR6, 23 defining key risks as potentially severe risks (Section 16.5). Key risks may refer to present or future 24 conditions, with a focus on the 21st century. Both mitigation and adaptation can moderate the extent or 25 severity of risks. The identification and evaluation of risks imply socio-cultural values, which may vary 26 across individuals, communities or cultures. 27 28 In line with chapter 16 of this report, this chapter uses a risk outcome perspective, i.e., the focus is on the 29 consequences related to risks, which potentially can result from different combinations of hazards, exposure 30 and vulnerabilities. There is limited literature with a focus on severe risks in the CSA region, and scant 31 studies specifically and explicitly considering risk drivers such as level of warming, level of exposure, 32 vulnerability and adaptation. 33 Do Not Cite, Quote or Distribute 12-44 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Criteria for identifying key risks for this chapter include the magnitude of the consequences, in particular the 2 number of people potentially affected; the severity of the negative effects of the risk (e.g., lives threatened, 3 major negative effect on livelihoods, well-being, or the economy); the importance of the affected system 4 (e.g., for vital ecosystem services, for large population groups); the irreversibility of either the process 5 leading to the risk or the consequences; and the potential to reduce the risk. 6 7 Several of the key risks identified for the CSA region align well with the overarching key risks assessed in 8 AR5 (Oppenheimer et al., 2014) and later in O'Neill et al. (2017), as well as with the representative key risks 9 assessed in Section 16.5 of this report. The identified key risks include KR1: risk of food insecurity due to 10 frequent and/or extreme droughts; KR2: risk to life and infrastructure due to floods and landslides; KR3: risk 11 of water insecurity; KR4: risk of severe health effects due to increasing epidemics (in particular vector-borne 12 diseases); KR5: systemic risks of surpassing infrastructure and public service systems KR6: risk of large- 13 scale changes and biome shifts in the Amazon; KR7: risk to coral reef ecosystems due to coral bleaching; 14 KR8: risks to coastal socio-ecological systems due to sea level rise, storm surges and coastal erosion (Table 15 12.6; Figure 12.11; Table SM12.5). 16 17 18 Table 12.6: Synthesis of key risks identified and assessed for the Central and South America region Consequence that would Associated changes in Associated changes in Associated changes in make the risk severe hazards exposure vulnerability 1. Risk of food insecurity due to frequent/extreme droughts Substantial decrease in yield More frequent and/or More people exposed to Reduced capacity of for key crops, disruption of longer drought periods. food insecurity due to farmers (especially small- food provision chains, Decrease in annual spatially more extensive scale) to adapt to reduced capacity or rainfall, severe decrease drought; high population changing climatic production of goods, reduced in rainfall at onset of growth rate (including conditions. Soil food security and increased rainy season. rural areas) and more degradation. Insufficient malnutrition. Desertification of population dependent on government support of semiarid regions. agricultural goods. adaptation measures, financial contributions, infrastructure, insurance, and research efforts. Inefficient water management. 2. Risk to life and infrastructure due to floods and landslides Death and severe health More frequent and severe More people exposed to Low income and marginal effects, disruption of critical storms and heavy floods and landslides due populations, low infrastructure and service precipitation events. to changing hazards, land- resilience of systems. Changing snow use and increased infrastructure and critical conditions and thawing population; occupation of service systems. Limited permafrost. Retreating more risk-prone areas such government support glaciers, formation of as flood plains and steep through insurance, glacier lakes, increased slopes. monitoring, early warning glacier lake outburst flood systems and recovery. hazard. 3. Risk of water insecurity Seasonal water availability Glacier shrinkage, snow Increase in population Unequal water change and decline due to cover change, more dependent on contribution consumption systems, glacier shrinkage, snow cover pronounced dry periods, of glacier/snow melt, failed water management Do Not Cite, Quote or Distribute 12-45 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report change, more pronounced dry precipitation and especially during drought and government periods and poor or failed circulation changes. conditions. Increased capacities, low water water management and demand from infrastructure efficiency, governance. intensification of growing urban areas. agriculture, mining, hydropower and urbanisation. 4. Risk of severe health effects due to increasing epidemics (in particular vector-borne diseases) Increased rate of epidemics of Higher temperatures Increased population Poor sanitation vector-borne diseases increase the geographical density and mobility conditions, particularly in (malaria, dengue, Zika, range of vectors, leading through urbanization low-income communities leishmaniasis) together with to expansion of climate results in high and for Indigenous diarrheal diseases. Severe suitable areas. transmission rate. Peoples. Insufficient health effects and damage to Increased population coverage of appropriate health systems in countries exposed to arboviruses due water provision and with low adaptive capacity to expansion of vectors, sewage systems. Low and where original endemicity including higher altitudes structural or economic is high and control status and latitudes. capacity to cope; poor. underfunding of health systems. Increase in infections can increase incidence of more severe forms of dengue. 5. Systemic risks of surpassing infrastructure and public service systems Breakdown of public service Higher frequency and More people and Increasing vulnerability systems, including magnitude of climate- infrastructure exposed to of public service and infrastructure and health related events (storms, climate/weather events. infrastructure systems. services due to cascading floods, landslides) Increase in population Insufficient disaster impacts of natural hazards and together with an increase exposed to arboviruses due management. Little epidemics, affecting a large in spatial and temporal to spatial expansion of improvement, part of the population. distribution of vectors. maintenance and pathogens/vectors for expansion of public malaria, dengue, Zika and health care systems. Low leishmaniasis. system resilience. 6. Risk of large-scale changes and biome shifts in the Amazon Transition from tropical forest More frequent, stronger Reduced availability of Strong dependence on into other biomes such as and persistent drought natural sources for local non-climatic drivers, in seasonal forest or savannah periods. Temperature people. Land use and land particular land-use through forest degradation increase and reduction in cover change (mining, change, deforestation, and deforestation. Risk of annual rainfall. deforestation). Loss of forest fire practices. Low shifting from carbon sink to biodiversity and ecosystem capacity to monitor and source. services. Health impacts control deforestation. from increased forest fires particularly for Indigenous Peoples. 7. Risk to coral reef ecosystems due to coral bleaching Do Not Cite, Quote or Distribute 12-46 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report Degradation and possible Ocean sea surface Continued exposure to Ecosystem highly death of the Mesoamerican temperature increase, increased atmospheric CO2 sensitive to water coral reef, the second largest lowered seawater pH and levels and sea surface temperature and pH reef in the world. Severe carbonate levels due to temperatures together with fluctuations. High levels damage to habitat for marine increased atmospheric destruction from coastal of negative human species, degrading coastal CO2 levels, leading to development, fishing interference with reefs protection and other ocean acidification and practices and tourism. including runoff and ecosystem services, decreased coral bleaching. pollution.. food security from fisheries, lack of income from tourism. 8. Risks to coastal socio-ecological systems due to sea level rise, storm surges and coastal erosion Coastal flooding and erosion High continuing Coastal population growth. Poor planning in coastal causing severe damage to trajectories of sea level Increased number of development and coastal population and rise. More intense and people, infrastructure and infrastructure, infrastructure. Loss of persistent coastal services (coastal tourism) disproportionate fisheries, reef degradation and flooding, salt water exposed; need of vulnerability and limited decline in coastal protection intrusion, coastal erosion. relocation of millions of adaptation options for due to increased storm surges people. rural communities and and waves. Salt water Indigenous Peoples, intrusion and land subsidence. increasing urbanisation in coastal cities. Large economic losses and unemployment from declining tourism. 1 2 3 Identification and assessment of key risks are informed by observed and projected impacts in the different 4 sub-regions of CSA (Section 12.3). Figure 12.10 shows the summary of different levels of observed and 5 future impacts per sub-region for different sectors, based on a detailed assessment of climate change impacts 6 on various systems and components for the respective sector (Figure 12.9). This assessment is consistent 7 with and complementary to the assessment in Section 12.3. A synthesis of these impacts (Figure 12.10) 8 indicates the following: Climate change has a major impact on observed and future decline of Andean 9 glaciers and snow (high confidence), and leads to degradation of permafrost and destabilization of related 10 landscapes (medium evidence, high agreement). Water quality is a major concern across the region but there 11 is limited evidence of impacts of climate change on water quality as well as on groundwater. Climate change 12 has had a high impact on terrestrial and freshwater ecosystems in the NWS, SES and SWS sub-regions, and a 13 medium impact in the other subregions but the level of confidence is varying across sub-region. Projections 14 indicate a strong impact of climate change on these ecosystems for the future (medium confidence: medium 15 evidence, high agreement). Many aspects and assets of ocean and coastal ecosystems (e.g., mangroves, coral 16 reefs, saltmarshes) were identified to be strongly impacted by climate change, both for observed and future 17 periods (high confidence) (Section 12.5.2; Figure 12.9). 18 19 Do Not Cite, Quote or Distribute 12-47 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 3 Figure 12.9: Observed and projected impacts for the subregions of Central and South America. Impacts are 4 distinguished for main sectors and for their corresponding systems (or components). Observed impacts refer to a time- 5 period of the last several decades. Projected impacts represent a synthesis across several emission and warming 6 scenarios, indicative of a time-period from mid- to end of the 21st century. For each system (e.g., coral reefs) it is 7 distinguished whether the impact of climate change is low, medium or high. The references underlying this assessment 8 can be found in SM12.4.1. 9 10 11 12 Figure 12.10: Synthesis of observed and projected impacts, distinguished for different sectors and each subregion of 13 Central and South America. Observed impacts refer to a time-period of the last several decades. Projected impacts 14 represent a synthesis across several emission and warming scenarios, indicative of a time-period from mid- to end of the 15 21st century. For each sector (e.g., health) it is distinguished whether the impact of climate change is low, medium or 16 high. The references underlying this assessment can be found in SM12.4.1 and the methodology to complete the 17 synthesis is found in SM12.4.2. Do Not Cite, Quote or Distribute 12-48 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 3 4 Figure 12.11: Synthesis of key risks for the Central and South America region. The base map indicates the mean 5 temperature change between the scenario SSP2 4.5 using CMIP6 model projections for 2081-2100, and a baseline 6 period of 1986-2005 (WGI AR6 Atlas, Gutiérrez et al., 2021). 7 8 9 In most sub-regions, crop, livestock, fisheries and food systems in general show medium to high impacts of 10 climate change over the observed period and similarly for the future of the 21st century (medium confidence: 11 medium evidence, high agreement). For some sub-regions, the available literature does not allow the 12 assessment of impacts on several human systems, including cities and infrastructure, health, poverty, 13 livelihoods, migration, conflict, Indigenous knowledge and local knowledge, especially for future time 14 periods. This points to important knowledge gaps about climate change impacts on human systems. 15 Indication of high impacts for several human systems and sub-regions points to the need to close these 16 knowledge gaps. 17 18 The assessment of key observed and projected impacts and risks shows that in the CSA region several 19 systems are already approaching critical thresholds under current warming levels, in particular glaciers in the 20 Andes and coral reefs in Central America (high confidence), and further ocean and coastal ecosystems in 21 virtually all sub-regions (medium confidence: medium evidence, high agreement). Some systems could cross 22 these thresholds with different levels of reversibility depending on the degrees of future warming, namely 23 glaciers in the Andes and coral reefs in Central America which will show partial but irreversible loss already 24 under low levels of warming (RCP2.6) (high confidence). The risk of large-scale ecological changes and 25 biome shifts of the Amazon forest, i.e., a transition from tropical forest into other biomes such as seasonal 26 forest or savannah, is now assessed with medium confidence, with the extent of the changes depending on the 27 level of future warming and non-climatic drivers (land-use change, deforestation, forest fire practices). 28 29 Systemic risks where critical infrastructure and public service system capacities are surpassed due to storms, 30 floods and epidemics, with cascading impacts through vulnerable systems and populations and economic 31 sectors, have the potential to affect large parts of the population and are therefore of major concern (medium 32 confidence: limited evidence, high agreement). The COVID-19 crisis has exposed the existing vulnerabilities 33 in important systems, in particular health systems and public service (Phillips et al., 2020). However, tipping 34 points in social systems are poorly understood (Bentley et al., 2014; Milkoreit et al., 2018), and there is Do Not Cite, Quote or Distribute 12-49 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 limited evidence to inform understanding about which level of compound climatic, environmental and socio- 2 economic stressors social systems withstand in CSA. 3 4 Overall, most key risks and their severity and extent are strongly driven and determined by the system's 5 exposure, vulnerability and adaptive capacity. In particular, the high vulnerability of large populations, 6 infrastructure and service systems such as health, food and energy production and supply are important 7 factors, along with high inequalities and poor governance, for creating and increasing key risks (high 8 confidence). Prevailing low levels of available information and understanding exacerbate the uncertainties 9 surrounding key risks, and hence pose limitations to adaptation. An example is Central America with high 10 levels of vulnerability and exposure but there is limited evidence and understanding on impacts and risks, 11 making this region susceptible to inappropriate adaptation to expected future climate change impacts. 12 13 14 12.5 Adaptation 15 16 Adaptation initiatives across the region have increased since AR5. National Communications (NC), 17 Nationally Determined Contributions (NDC) and National Adaptation Plans (NAP) (https://unfccc.int) 18 recently published are providing guidance for adaptation in CSA. There is also a diversity of non- 19 governmental adaptation initiatives, both at the national and sub-national levels. In this context, this section 20 assesses, through a sectoral approach, the main challenges, opportunities, trends and initiatives to adapt to 21 climate change in the region. 22 23 12.5.1 Terrestrial and Freshwater Ecosystems and their Services 24 25 CSA is one of the most biodiverse regions in the World, hosting unique socio ecosystems that will be 26 strongly impacted by climate change (high confidence) (Section 12.3; Cross-Chapter Paper 1; CAF, 2014; 27 Camacho Guerreiro et al., 2016; IPBES, 2018a; Li et al., 2018; Retsa et al., 2020) . Warming has generated 28 extreme heat events in many parts of CSA (IPCC, 2019a) that, together with droughts and floods, will 29 seriously affect the integrity of terrestrial and freshwater ecosystems in the entire region (Section 12.3; CAF, 30 2014) . A reduction in net primary productivity in tropical forests and glacier retreat in the Andes, for 31 example, are expected to cause significant negative socioecological impacts (Feldpausch et al., 2016; Lyra et 32 al., 2017; Cuesta et al., 2019) (see Case Study, 12.7.1). Biodiversity-rich spots in the region are well assessed 33 in the literature as compared to other regions of the World, especially for the Atlantic Forest, Mesoamerica 34 and Cerrado (Cross-Chapter Paper 1.2.2; Manes et al., 2021) . Up to 85% of of evaluated natural systems 35 (species, habitats and communities) in the literature for biodiversity-rich spots since AR5 were projected to 36 be negative impacted by climate change (high confidence), with 26% of projections predicting species 37 extinctions (Cross-Chapter Paper 1.2.2; Manes et al., 2021) . Indigenous knowledge and local knowledge 38 play an important role in adaptation and are vital components of many socioecological systems, while also 39 being threatened by climate change (high confidence) (Box 7.1; Valdivia et al., 2010; Tengö et al., 2014; 40 Mistry et al., 2016; Harvey et al., 2017; Diamond and Ansharyani, 2018; Camico et al., 2021) . 41 42 12.5.1.1 Challenges and opportunities 43 44 The conversion of natural ecosystems to agriculture, pasture and other land uses in CSA has been identified 45 as a major challenge to climate change adaptation in the region (high confidence) (Scarano et al., 2018; 46 IPCC, 2019a). In the last three decades, South America has been a significant contributor of the growth of 47 agricultural production worldwide (OECD/Food and Agriculture Organization of the United Nations, 2015), 48 driven partly by increased international demand for commodities, especially soybeans and meat (IPCC, 49 2019a). Between 2001 and 2015 about 65% of all forest disturbance in the region was associated with 50 commodity-driven deforestation (Curtis et al., 2018). High rates of native vegetation conversion in 51 Argentina, Bolivia, Brazil, Colombia, Ecuador, Paraguay and Peru threaten important ecosystems (Amazon, 52 Cerrado, Chacos and Llanos savannas, Atlantic rainforest, Caatinga and Yungas) (Graesser et al., 2015; 53 FAO, 2016c). Almost 2/3 of soy consumed in EU+ comes from Brazil, Argentina and Paraguay (IDH, 2020), 54 increasing conversion risk in the Amazon, Cerrado, and Gran Chaco. Despite growing commodities 55 production traceability, in 2018 only 19% of the soybean meal consumed in EU+ was certified deforestation- 56 free, and 38% compliant with the FEFAC Soy Sourcing Guidelines (IDH, 2020), which is a great challenge 57 at the international level (Negra et al., 2014; Curtis et al., 2018; Lambin et al., 2018; IDH, 2020). Do Not Cite, Quote or Distribute 12-50 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 Investing in actions aimed at protection, restoration and sustainable use of biodiversity and ecosystems is a 3 good approach for maintaining critical ecosystem services, and is part of a common strategy for adaptation, 4 mitigation and disaster risk reduction in the region (high confidence) (Kabisch et al., 2016; Scarano et al., 5 2018). These strategies also meet the forest and water conservation international agendas, optimizing 6 resources and solutions (Strassburg et al., 2019). Global conservation and sustainable development 7 commitments, such as the Aichi Targets (CDB), Sustainable Development Goals (UN), the Nationally 8 Determined Contribution (NDC) under the Paris Agreement, and the New York Declaration on Forests 9 strongly rely on nature-based solutions (NbS) to achieve their objectives (Brancalion et al., 2019) (Figure 10 12.12). The COVID­19 outbreak also brought attention to the need for preserving tropical forests as a mean 11 to prevent spill over of viruses from wildlife to humans, with concerns over that risk in the Amazon (Allen et 12 al., 2017b; Dobson et al., 2020; IPBES, 2020; Ferreira et al., 2021). These represent an important 13 opportunity for Ecosystem Based Adaptation (EbA) to be at the core of NbS for climate change, access 14 finance and promote climate resilient development pathways in CSA. 15 16 The Declaration on Protected Areas and Climate Change, presented by 18 CSA countries during the 17 UNFCCC COP21, highlights the fundamental role of protected areas in providing the "green infrastructure" 18 needed for implementing climate change mitigation and adaptation, and safeguard the provision of essential 19 ecosystem services and the livelihoods of Indigenous Peoples and local communities (Gross et al., 2016). 20 Protected Areas systems in CSA are underfunded (very high confidence). Latin American (including 21 Mexico) governments allocate just about 1% of national environmental budgets on protected areas (about 22 USD 1.18 ha-1 on average). This figure only covers 54% of their basic needs, resulting in insufficient 23 management. The financing gap to achieve optimal needs for protected areas in CSA is approximately USD 24 700 million yr-1 (Bovarnick et al., 2010). This seriously compromises the management and delivery capacity 25 of protected areas for climate change adaptation, and preparedness for ongoing ecological transformation 26 (van Kerkhoff et al., 2019). Furthermore, in order to become a relevant mechanism for resilience, protected 27 areas need to be managed for this purpose (Mansourian et al., 2009). About 40% of protected areas in Latin 28 America and Caribbean (including Mexico), have management effectiveness evaluations being undertaken 29 (UNEP-WCMC and IUCN, 2020a). This is hardly representative of Aichi's Goal 11, although far better than 30 the 11% global average. Collaborations with the Indigenous Peoples and local communities are also an 31 important issue to consolidate protected areas (Gross et al., 2016). In addition to protected areas as solutions 32 for climate change adaptation and mitigation, there is also a need to protect or restore ecosystems outside the 33 protected areas, as illustrated by the Mesoamerican Biological Corridor (Imbach et al., 2013). 34 35 Despite some local and specific assessments (e.g., Warner (2016)), there is a significant gap on identifying 36 barriers to adaptation or maladaptation in the region (Dow et al., 2013). In their National Communications 37 (NC), Nationally Determined Contributions (NDC) and/or National Adaptation Plans (NAP) 38 (https://unfccc.int), most countries identified inadequate financing and access to technology as barriers for 39 adaptation relevant to terrestrial and freshwater socio-ecosystems (high confidence). Insufficient institutional 40 coordination is also frequently mentioned (Rangecroft et al., 2013; Cameron et al., 2015). These limitations 41 could be partially addressed through multilateral cooperation, incorporation of synergies from the local to the 42 national scales, local empowerment, and poverty alleviation (Rangecroft et al., 2013; Harvey et al., 2017; 43 Murcia et al., 2017; Calispa, 2018; Chain-Guadarrama et al., 2018). 44 45 12.5.1.2 Governance and financing 46 47 All CSA countries have formulated policies that include measures relevant for socio-ecosystem adaptation in 48 their NCs, NDCs and NAPs (https://unfccc.int), with an emphasis on protection and restoration of water and 49 forests (high confidence). Existing proposed measures, instruments and programs, however, do not yet 50 reflect the vision needed to integrate the ecosystem and human dimensions of vulnerability. The 51 administration coordination and the progress in adaptive ecosystem management are incipient, due in part to 52 the lack of stable financial resources and scientific, Indigenous knowledge and local knowledge (IK and LK) 53 about adapting ecosystems to climate change (Bustamante et al., 2020). Brazil was an exception, showing 54 dramatic policy-driven reduction in deforestation in the Amazon between 2004­2012, with a concomitant 55 70% increase in soy production, the most profitable Amazon crop (Hansen et al., 2013; Nepstad et al., 2014). 56 Policies included territorial planning (protected areas, Indigenous territories and land tenure), satellite 57 monitoring, market and credit restrictions on high-deforesting municipalities, plus some incentives to small Do Not Cite, Quote or Distribute 12-51 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 farmers (Boucher et al., 2013; Hansen et al., 2013; Nepstad et al., 2014; Castelo, 2015; Cunha et al., 2016a). 2 It is important to highlight the important role of Indigenous territories, in addition to protected areas, in 3 forest conservation in the Amazon (high evidence, medium agreement) (Schwartzman et al., 2013; Barber et 4 al., 2014; Nepstad et al., 2014; Walker et al., 2014b). These policies were partially funded by results-based 5 compensation through the Amazon Fund. Since 2012, however, policies and institutions have weakened, and 6 Amazon deforestation rates started to rise (Carvalho et al., 2019), sharpening in recent years (Silva Junior et 7 al., 2021). Conservation incentives, a new complementary and allegedly cost-effective approach, is 8 increasingly being implemented in the region (Magrin et al., 2014). They include payment for ecosystem 9 services, REDD+, environmental certification and conservation easements, but remain controversial, and 10 more research is needed on their effectiveness, possible negative side effects, participatory management 11 systems and collective decision-making processes (Larson and Petkova, 2011; Locatelli et al., 2011; Pinho et 12 al., 2014; Strassburg et al., 2014; Mistry et al., 2016; Gebara and Agrawal, 2017; Scarano et al., 2018; 13 Ruggiero et al., 2019; To and Dressler, 2019; Vallet et al., 2019). 14 15 12.5.1.3 Adaptation options to avert and reduce key risks on terrestrial and freshwater ecosystems 16 17 Research, monitoring systems and other initiatives for knowledge management are promoted in the region on 18 terrestrial and freshwater socio-ecosystem adaptation (high confidence) (NCs, NDCs and NAPs, 19 https://unfccc.int). In Chile, for example, the Eco-social Observatory of Climate Change Effects for High 20 Altitude Wetlands of Tarapacá has been collecting information on physical, biological and social variables 21 since 2013 (Uribe Rivera et al., 2017). Other examples in the Andes are the GLORIA-Andes network 22 (Cuesta et al., 2017a), the Andean Forest Network (Malizia et al., 2020) and the Initiative of Hydrological 23 Monitoring in the Andes (IMHEA), with measures to optimize watershed management and protection, and 24 reduce the risk of water insecurity (Correa et al., 2020). 25 26 Poverty is a driver of climate change risk, while sustainable use of ecosystems fosters adaptation (Kasecker 27 et al., 2018) (high confidence). Most of 398 "Ecosystem-based Adaptation hotspots" identified in Brazil on 28 this premise are located in some of the most vulnerable ecosystems to climate change (Kasecker et al., 2018). 29 Although conservation and restoration is reported as effective to reduce risk (medium confidence: medium 30 evidence, high agreement) (Anderson et al., 2010; Borsdorf et al., 2013; Keenan, 2015; Pires et al., 2017; 31 Ramalho et al., 2021), their effectiveness depends on the integration of conservation actions with 32 enhancement of local socioeconomic conditions (medium confidence: medium evidence, high agreement) 33 (Scarano and Ceotto, 2015; Pires et al., 2017; Kasecker et al., 2018; de Siqueira et al., 2021; Vale et al., 34 2021). 35 36 Since AR5, there has been an increase in the number of adaptation measures through natural resources and 37 ecosystem services management. The main approaches are EbA and Community-based Adaptation (CbA) 38 (high confidence) (NCs, NDCs and NAPs, https://unfccc.int). IK/LK can be very detailed and usually relates 39 to people's priorities identified by collective decision-making (Box 7.1; (Hurlbert et al., 2019, SRCCL 40 Section 7.6.4); SRCCL Cross-Chapter Box ILK in Chapter 13; (de Coninck et al., 2018, SR1.5 Section 41 4.3.5.5). In Manaus, central Amazon, fishermen perceive reductions on fish size, diversity and capture levels 42 caused by droughts; while recognizing that floods hinders access to fishing grounds (Keenan, 2015; 43 Camacho Guerreiro et al., 2016). In the Amazon floodplains, small-scale fisher and farmer's communities 44 incorporate their knowledge on natural hydrologic and ecological processes into management systems that 45 reduce climate change risk and impacts (Oviedo et al., 2016). Smallholder grain farmers in Guatemala and 46 Honduras implement EbA practices based on local knowledge (e.g., live fences, home gardens, shade trees in 47 coffee plantations, dispersed trees in corn fields and other food insecurity risk reduction practices) (Harvey et 48 al., 2017; Chain-Guadarrama et al., 2018). There is, therefore, a great potential for terrestrial and freshwater 49 ecosystem adaptation to climate change in CSA, provided that the right incentives and sociocultural 50 protective measures are in place (high confidence) (Section 12.5.10.4; Table SM12.7). 51 52 Disarticulation between policy and implementation is a common problem. Ecuadorian climate public policy 53 points towards a CbA approach, but it is often downsized in the implementation (Calispa, 2018). Important 54 adaptation actions have been undertaken in Argentina, Bolivia, Brazil, Chile, Colombia, Ecuador, El 55 Salvador, Paraguay, Peru and Uruguay; both in policymaking and institutional arrangements, but they tend to 56 be poorly coordinated with policies on development, land planning and other sectoral policies (Ryan, 2012). 57 Some type of community participation mechanisms is present in most country strategies, but their levels of Do Not Cite, Quote or Distribute 12-52 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 implementation vary considerably (medium confidence: medium evidence, high agreement) (Ryan, 2012; 2 Pires et al., 2017; Calispa, 2018). 3 4 There is an ecosystem bias in adaptation priorities for research and implementation, hindering the 5 development of comprehensive adaptation programs. Most scientific research on adaptation in Peru focuses 6 on the highlands and coastal regions while mitigation research focuses on forests (Chazarin et al., 2014). 7 Combined adaptation and mitigation strategies can produce positive results, but they are often disconnected 8 (Locatelli et al., 2015). Most reviewed cases in agriculture and forestry in Latin America (84% of 274 cases) 9 reported positive synergies between adaptation and mitigation. Nevertheless, research on Latin American 10 forests tend to focus on mitigation, while studies on agriculture are usually oriented towards adaptation (high 11 confidence) (Locatelli et al., 2015; Locatelli et al., 2017). 12 13 Rural communities in the Cusco Region, Peru, ground their ability to adapt to climate change on four 14 cultural values, known in Quechua as ayni (reciprocity), ayllu (collectiveness), yanantin (equilibrium) and 15 chanincha (solidarity), but policies oriented towards "modernization" undermine these traditional 16 mechanisms. Adaptation strategies could benefit from integrating these and other insights from traditional 17 cultures, fostering risk reduction and transformational adaptation towards intrinsically sustainable systems 18 (medium confidence: medium evidence, high agreement) (Walshe and Argumedo, 2016). 19 20 Protected areas have become an important component as enablers of national climate change adaptation 21 strategies. They increase ecosystem's adaptive potential, reducing climate risk and delivering numerous 22 ecosystem services, sustainable development benefits while playing an important role in climate change 23 mitigation (high confidence) (Mackey et al., 2008; Dudley et al., 2010; Gross et al., 2016; Bebber and Butt, 24 2017; Dinerstein et al., 2019; IPCC, 2019a). CSA already has a greater percentage of land (24.1%) under 25 protected status than the world average (14.7%) (UNEP-WCMC and IUCN, 2020b). Some countries, 26 including Belize, Bolivia, Brazil, Guatemala, Nicaragua and Venezuela already met or surpassed the 30% 27 CDB and IUCN goal (Dinerstein et al., 2019), and others like Costa Rica and Honduras are very close to 28 doing so. In some cases, the establishment of protected areas not accompanied by collective decision-making 29 processes has displaced local people or denied them access to natural resources, increasing their vulnerability 30 to climate change (Brockington and Wilkie, 2015). 31 32 In addition to better managing and expanding protected areas networks, Other Effective Area-based 33 Conservation Measures (OECMs), recently defined by the Parties to the Convention on Biological Diversity 34 (Dudley et al., 2018), could also enhance ecosystem resilience (low confidence). Private Protected Areas in 35 the mountain regions of the Americas (e.g., Andes), play an important role in closing the gaps in fragmented 36 biomes and expanding protection in underrepresented areas (Hora et al., 2018). In Brazil, there is also a huge 37 potential for conservation and sustainable management in private areas, as roughly 53% of the country's 38 native vegetation is within private land (Lapola et al., 2014; Soares-Filho et al., 2014). 39 40 Large-scale restoration is also seen as pivotal to limiting both climate change (IPCC, 2019a) and species 41 extinction (IPBES, 2018a) (very high confidence). A new multi-criteria approach for optimizing multiple 42 restoration outcomes (for biodiversity, climate change mitigation, and cost), for example, indicate that South 43 America has the greatest extension of converted lands, evenly distributed in the top 50% of global priorities 44 (Strassburg et al., 2020). 45 46 12.5.2 Ocean and Coastal Ecosystems and their Services 47 48 Ocean and coastal ecosystems provide suitable habitats to a high number of species that support important 49 local fisheries, the tourism sector and the economy of the region (high confidence) (Section 3.5; Table 3.9; 50 González and Holtmann-Ahumada, 2017; Venerus and Cedrola, 2017; CEPAL, 2018; Carvache-Franco et 51 al., 2019; SROCC Section 5.4 Bindoff et al., 2019). There is high confidence that CSA ocean and coastal 52 ecosystems are already impacted by climate change (Figure 12.9, 12.10; Table SM12.3; Section 3.4; , 53 Section 5.4 in SROCC, Bindoff et al., 2019), and highly sensitive to non-climate stressors (Figure 12.8; 54 Table SM12.3; Section 3.4). Projections for CSA ocean and coastal ecosystems alert about significant and 55 negative impacts (high confidence) which include major loss of ecosystem structure and functionality, 56 changes in the distributional range of several species and ecosystems, major mortality rates, and increasing Do Not Cite, Quote or Distribute 12-53 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 number of coral bleaching events (Figure 12.9; Figure 12.10; Table SM12.3; Section 3.4; SROCC Sections 2 5.3, 5.4, Bindoff et al., 2019). 3 4 CSA subregions are highly dependent on ocean and coastal ecosystems, and thus vulnerable to climate 5 change (FAO, 2018). Fisheries and aquaculture contribute significantly to food security and livelihoods by 6 creating employment (more than two million people), income and economic growth for the region (Section 7 3.5; FAO, 2018) (). More than 45% of the total fisheries in CSA are based on marine products 8 (CEPALSTAT, 2019). Peru, Chile, Argentina and Ecuador are among the 15 countries with the largest 9 marine capture production worldwide (Gutiérrez et al., 2016a; FAO, 2018; Vannuccini et al., 2018), while 10 more than 90% of the hydrological resources produced by aquaculture in CSA have a marine origin 11 (CEPALSTAT, 2019). There is high confidence about important current and future impacts of climate 12 change hazards in marine resources subjected to fisheries, however there is low evidence about the impacts 13 on regional economies (Figure 12.9, 12.10; Table SM12.3). 14 15 12.5.2.1 Adaptation measures and strategies applied on oceans and coasts of CSA 16 17 Similar to those pointed by WGII AR5 Chapter 27 (Magrin et al., 2014) and Chapter 3 (Section 3.5; Section 18 3.6.2; Box SLR in Chapter 3), adaptation strategies in ocean and coastal ecosystems in CSA are still focused 19 on the ecosystem protection and restoration, and the sustainable use of marine resources (high confidence). 20 There is low evidence about how coastal urban areas and touristic settlements of CSA countries are adapting 21 to SLR and extreme events (Calil et al., 2017; Villamizar et al., 2017). Some of this strategies include 22 planned relocation (Dannenberg et al., 2019) and the use of grey infrastructures as seawalls and bulkheads 23 (Silva et al., 2014; Isla et al., 2018) . 24 25 There is medium confidence that Ecosystem-based Adaptation (EbA) is the main strategy used in CSA coral 26 reefs ecosystems. The set of strategies applied include the protection, restoration (e.g., coral gardening, larval 27 propagation), and conservation of coral reefs areas through the application of the spatial ocean zoning 28 schemes such as Marine Protected Areas (MPAs), marine managed areas (MMAs), National Parks, Wildlife 29 Refuges, Special Zones of Marine Protection, Special Management Zones, Responsible Fishing Areas, and 30 the establishment of management plans with some level of participatory processes. These strategies are 31 complemented with actions that promote the development of research and education programs, recreational 32 and cultural activities, the use of community-based approaches, and the creation of national specific laws 33 (Graham, 2017) and the adhesion of international treaties (e.g., Convention on International Trade in 34 Endangered Species of Wild Fauna and Flora (CITIES), AGENDA 21, United Nations Convention on the 35 Law of the Sea (UNCLOS), Ramsar Convention on Wetlands of International Importance Especially as 36 Waterfowl Habitat) (Cruz-Garcia and Peters, 2015; Gopal et al., 2015; Graham, 2017; Bayraktarov et al., 37 2020). 38 39 Adaptation measures in mangroves ecosystems are mainly focused on the application of EbA strategies (high 40 confidence). This measures include the application of restoration programs, the creation of management 41 plans (which also have significant co-benefits with mitigation (Section 3.6.2.1), and the establishment of 42 coastal protected areas, followed by the development of research activities, the creation of specific mangrove 43 policies through new laws and resolutions (e.g., Colombia) (Cvitanovic et al., 2014; Krause, 2014; Blanco- 44 Libreros and Estrada-Urrea, 2015; Carter et al., 2015; Estrada et al., 2015; Ferreira and Lacerda, 2016; 45 Oliveira-Filho et al., 2016; Rodríguez-Rodríguez et al., 2016; Alvarado et al., 2017; Álvarez-León and 46 Álvarez Puerto, 2017; Baptiste et al., 2017; Borges et al., 2017; Jaramillo et al., 2018; Salazar et al., 2018; 47 Armenteras et al., 2019; Blanco-Libreros and Álvarez-León, 2019; Maretti et al., 2019; Ellison et al., 2020) 48 49 The use of territorial planning tools, the promotion of sustainable resource exploitation, the adherence to 50 certification schemes, and the implementation of management instruments such as Ecosystem-based 51 Management (EbM) followed by the use of an integrated coastal zone management, coastal marine spatial 52 planning, capacity building, ecological risk assessments have been the mains strategies used to ensure the 53 sustainability of marine resources subjected to fisheries across EEZs of CSA (high confidence) (Hellebrandt 54 et al., 2014; Gelcich et al., 2015; Singh-Renton and McIvor, 2015; Gutiérrez et al., 2016a; Karlsson and 55 Bryceson, 2016; Oyanedel et al., 2016; Debels et al., 2017; Isaac and Ferrari, 2017; Mariano Gutiérrez et al., 56 2017; Barragán and Lazo, 2018; Bertrand et al., 2018; Lluch-Cota et al., 2018; Guerrero-Gatica et al., 2020). Do Not Cite, Quote or Distribute 12-54 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Other strategies include the application of local regulations (e.g., closed seasons) (Fontoura et al., 2016), and 2 the use of participative instances (Hellebrandt et al., 2014; Arroyo Mina et al., 2016; Matera, 2016). 3 4 12.5.2.2 Adaptation success in ocean and coastal ecosystems of CSA 5 6 There is low evidence about how the strategies and actions taken and implemented in ocean and coastal 7 systems of CSA have contributed to advance in the protection and conservation of ocean and coastal 8 ecosystems. However, some important advances are visible in Colombian Pacific areas with coral reefs (new 9 conservation plans, research monitoring and conservation practices) (low confidence) (Cruz-Garcia and 10 Peters, 2015; Alvarado et al., 2017; Bayraktarov et al., 2020). In Panama, actions taken have allowed the 11 protection of a high number of marine areas with coral reefs, as well as the incorporation of management 12 approaches that include several sectors such as fisheries, tourism, coral protection and coral conservation 13 (low confidence) (Alvarado et al., 2017). In the case of Costa Rica, 80% of coral habitats are located inside 14 of MPAs, multiple research coral-related activities have been performed, and several training activities have 15 favoured the engagement of the local community in their protection against climate and non-climate hazards 16 (low confidence) (Alvarado et al., 2017). 17 18 There is low evidence of how the incorporation of mangroves as Ramsar sites, the reforms of legislations 19 (e.g., fines and stronger regulations), and the creation of reserves and private protection initiatives (e.g., 20 Belize Association of Private Protected Areas BAPPA), and capacity-building projects or new educational 21 programs have promoted the protection of mangroves in CSA countries such as Honduras, Guatemala and 22 Belize (Cvitanovic et al., 2014; Carter et al., 2015; Ellison et al., 2020). In Brazil, between 75­84% of 23 mangroves are under some level of protection which has improved the forest structures, and multiple 24 research programs (e.g., Mangrove Dynamics and Management, MADAM, and `GEF-Mangle') have been 25 developed (medium confidence) (Krause, 2014; Medeiros et al., 2014; Estrada et al., 2015; Ferreira and 26 Lacerda, 2016; Oliveira-Filho et al., 2016; Borges et al., 2017; Maretti et al., 2019; Strassburg et al., 2019). 27 In Colombia, research projects (e.g., Mangroves of Colombia Projects, MCP), the installation of a 28 geographic information system for mangroves (e.g., SIGMA Sistema de Información para la Gestión de los 29 Manglares en Colombia), surveillance monitoring plans (e.g., EGRETTA Herramientas para el Control y 30 Vigilancia de los Manglares), and the establishment of protected areas have contributed to decrease loss of 31 the mangrove forest (high confidence) (Blanco-Libreros and Estrada-Urrea, 2015; Rodríguez-Rodríguez et 32 al., 2016; Álvarez-León and Álvarez Puerto, 2017; Baptiste et al., 2017; Jaramillo et al., 2018; Salazar et al., 33 2018; Armenteras et al., 2019; Blanco-Libreros and Álvarez-León, 2019). 34 35 There is low evidence whether the establishment of MPAs and the creation of legal instruments have allowed 36 the development of new research activities have increased the environmental awareness, decreased the illegal 37 extraction, and improved the local coordination which have promoted the sustainable use of marine 38 resources, and improved the community-government cooperation in marine ecosystems (Alvarado et al., 39 2017). The experience in countries like Chile demonstrates the importance of implementing robust 40 management plans that guarantee the protection objectives and the sustainability through the implementation 41 of EbA measures such as MPAs (Petit et al., 2018). 42 43 There is low confidence about how measures adopted are ensuring the sustainability of marine resources 44 subjected to fisheries. In Peru, the industrial fishery follows an adaptive management approach (i.e., stock 45 assessments, catch limits), while in Chile, the small-scale fishery of benthic-demersal resources is managed 46 through the granting of exclusive territorial use rights (called TURFS) with established quotas defined by the 47 central authority (Bertrand et al., 2018). In addition, MPAs in Chile are playing a key role in climate change 48 adaptation for fisheries (medium confidence) (Gelcich et al., 2015; Petit et al., 2018), and an increasing 49 amount of funds have been invested in initiatives to reduce the vulnerability of fishery and aquaculture 50 sectors to climate change (OECD, 2017). Since 2016, Argentina has been developing a strategy to implement 51 EbM on fisheries with support from the Global Environment Facility program (GEF). Also, Argentina and 52 Chile, are promoting the local consumption of seafood and the certification of its fishery products (OECD, 53 2017), while Brazil and Chile have advanced in their actions to climate change through the development of 54 new research studies and methodologies incorporating research institutions (Nagy et al., 2015). Uruguay is 55 incorporating stakeholders in their climate change adaptation strategies (low confidence) (Nagy et al., 2015), 56 while Colombia is supporting the capacity building of fishers promoting livelihood diversification to 57 increase the resilience of the sector (medium confidence: medium evidence, high agreement) (Hellebrandt et Do Not Cite, Quote or Distribute 12-55 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 al., 2014; Arroyo Mina et al., 2016; Matera, 2016). Chile and Peru have showed certain advances in the 2 development of guidelines for the management of the coast line and the implementation of the EbM which 3 has favoured the collaboration of diverse and multiple stakeholders (fishers, academics, municipal 4 institutions), the development of outreach and educational activities, and the creation of networks, and the 5 interest of other fishery communities to implement EbM (medium confidence: medium evidence, high 6 agreement) (Hellebrandt et al., 2014; Gelcich et al., 2015; Gutiérrez et al., 2016a; Oyanedel et al., 2016; 7 Guerrero-Gatica et al., 2020). In countries like Peru and Chile, there is an increasing presence of 8 intergovernmental and international cooperation agencies, and new funding (e.g., GEF), and projects (Inter- 9 American Development, SPINCAM) related to change adaptation for the fishery sector (medium confidence: 10 medium evidence, high agreement) (Galarza and Kámiche, 2015; Barragán and Lazo, 2018). 11 12 12.5.2.3 National climate change commitments for ocean and coasts 13 14 Beyond the protection, conservation and climate change adaptation strategies implemented on CSA ocean 15 and coastal areas and their ecosystems, a high number of adaptation goals to face climate change impacts on 16 ocean and coastal ecosystems and their services are incorporated in most of the national climate change 17 adaptation commitments of CSA countries (Table 12.7). 18 19 20 Table 12.7: National plans with adaptation goals for ocean and coasts in CSA. CSA country Adaptation Initiatives Year Argentina Plan Nacional de Adaptación y Mitigación al Cambio Climático1 2019 Brazil National Adaptation Plan to Climate Change (Volume 1); General Strategies2 2016 National Adaptation Plan to Climate Change (Volume 2); Sectoral and thematic 2016 strategies3 Chile Plan Nacional de Adaptación al Cambio Climático4 2014 Plan Sectorial de Adaptación al Cambio Climático en Biodiversidad5 2014 Plan Sectorial de Adaptación al Cambio Climático en Pesca y Acuicultura6 2015 Plan de Adaptación y Mitigación de los Servicios de Infraestructura al Cambio 2017 Climático7 Plan de Adaptación al Cambio Climático Sector Salud8 2017 Colombia Plan Nacional de Adaptación al Cambio Climático9 2016 Costa Rica Política Nacional de Adaptación al Cambio Climático10 2018 Ecuador Plan Nacional de Cambio Climático11 2015 El Salvador Plan Nacional de Cambio Climático12 2015 Guatemala Plan de Acción Nacional de Cambio Climático13 2018 Guyana Política de Adaptación y Plan de Implementación14 2001 Honduras Plan Nacional de Adaptación al Cambio15 2018 Nicaragua Plan de Adaptación a la Variabilidad y el Cambio Climático en el Sector 2013 Agropecuario, Forestal y Pesca16 Peru Plan Nacional de Adaptación al Cambio Climático del Peru17 2021 Suriname Suriname National Adaptation Plan18 2019 Uruguay Plan Nacional de Respuesta al Cambio Climático19 2010 Belize Not Available 2019 Panamá Not Available Venezuela Not Available References: 1(Ministerio de Ambiente y Desarrollo Sostenible de la República de Argentina, 2019) 2(Ministry of Environment of Brazil, 2016a) 3(Ministry of Environment of Brazil, 2016b) 4(Ministerio de Medio Ambiente de Chile, 2014b) 5(Ministerio de Medio Ambiente de Chile, 2014a) 6(Ministerio de Economía Fomento y Turismo de Chile, 2015) 7(Ministerio de Medio Ambiente de Chile, 2017) 8(Ministerio de Salud de Chile, 2017) 9(Ministerio de Ambiente y Desarrollo Sostenible de Colombia, 2016) 10(Ministerio de Ambiente y Energía de la República de Costa Rica, 2018) 11(Gobierno Nacional de la República del Ecuador, 2015) 12(Ministerio de Medio Ambiente y Recursos Naturales de El Salvador, 2015) 13(Consejo Nacional de Cambio Climático y la Secretaría de Planificación y Programación de la Presidencia de Guatemala, 2018) 14(National Ozone Action Unit of Guyana, 2016) 15(Secretaría de Recursos Naturales y Ambiente del Gobierno de la República de Honduras, 2018) 16(Ministerio Agropecuario y Forestal de Nicaragua, 2013) 17(Ministerio del Ambiente Gobierno del Perú, 2021) Do Not Cite, Quote or Distribute 12-56 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 18(Government of Suriname, 2019) 19(Ministerio de Vivienda Ordenamiento Territorial y Medio Ambiente de la República de Uruguay, 2010) 1 2 3 Current goals in national and sectoral adaptation plans attempt to promote research and monitoring (e.g., new 4 research actions, modelling, knowledge management), the development of new legislation tools and policies 5 (e.g., inter-institutional and territorial coordination, improvement of public policies), the conservation of 6 ocean and coastal ecosystems and their biodiversity (e.g., new MPAs establishment, protection tools), the 7 management of climate risks (e.g., alert systems), the management of productive activities (e.g., 8 diversification of resources), the promotion of the construction of new infrastructure and technology (e.g., 9 grey-green infrastructure - GGI), the creation of new financial tools (e.g., insurances), the improvement of 10 the capacity building (e.g., education, awareness), the management of water and residues (e.g., sewages and 11 freshwater availability), the social inclusion (e.g., strategies to support vulnerable sectors, gender inclusion), 12 and the incorporation of traditional practices (e.g., restoring traditional practices including Indigenous 13 knowledge). However, the amount and the type of adaptation goals per country differ enormously among 14 countries (Figure 12.12). 15 16 17 18 19 Figure 12.12: Type and amount of adaptation goals identified in National Adaptation Plans for ocean and coastal 20 systems of CSA countries. 21 22 23 12.5.2.4 Limits and barriers for adaptation in ocean and coastal ecosystems 24 25 Although current national adaptation plans and many other actions and strategies are focused on improving 26 the conservation and restoration of ocean and coastal ecosystems, as well as, the suitability of marine 27 resources along CSA, these measures are still not able to reduce the vulnerability and sensitivity of these 28 ecosystems to climate change hazards (high confidence) (Figure 12.6; Table SM12.3; Leal Filho, 2018; Nagy 29 et al., 2019) . There is high confidence that sandy beaches ecosystems of CSA countries show an important 30 loss of dunes as a consequence of the construction of infrastructures which have generate an interruption of 31 the natural dynamic of beaches decreasing the protection to tides, waves, extreme events or tsunamis (high 32 confidence) (Amaral et al., 2016; Bernardino et al., 2016; González and Holtmann-Ahumada, 2017; 33 Obraczka et al., 2017). Also, adaptation measures to cope with SLR and coastal extreme events sometimes Do Not Cite, Quote or Distribute 12-57 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 fail as they exacerbate coastal erosion and damage (medium confidence: medium evidence, high agreement) 2 (Spalding et al., 2014; Lins-de-Barros and Parente-Ribeiro, 2018). There is medium evidence but high 3 agreement that the most barriers limiting the success of adaptation strategies in ocean and coastal systems in 4 CSA are due to the lack of coordination (e.g., absence of participatory processes, overlapping among fishing 5 and protection activities), the lack of knowledge (e.g., poor monitoring, poor control and surveillance, no 6 long-term studies), lack of adequate metrics for evaluating adaptation actions informing decision-makers 7 hinder the continuity and adjustment of measures, weak governance (e.g., perverse incentives, resource 8 overexploitation, conflicts), lack of financial resources and long-term commitments (e.g., crisis, lack of 9 budgets, market fluctuations), weak policies, cultural constraints, poverty, low flexibility, lack of awareness 10 of climate risks, and lack of engagement by stakeholders (Leal Filho, 2018; Nagy et al., 2019; Moreno et al., 11 2020b; Aburto et al., 2021). 12 13 Some important limits and barriers have been detected for productive systems such as fisheries and tourism 14 in CSA (medium confidence: medium evidence, high agreement). Brazilian major fisheries management do 15 not follow an ecosystem approach, although some small-scale fisheries apply a precautionary approach 16 (Singh-Renton and McIvor, 2015). The management of Peruvian artisanal (medium and small-scale 17 fisheries) are minimal with an important lack of regulations, control, and management actions (Bertrand et 18 al., 2018). In Argentina, marine recreational fisheries have been largely unregulated with a lack of 19 monitoring programs which have contributed to the overexploitation of some key coastal stocks (Venerus 20 and Cedrola, 2017). Moreover, the participation of women fishers in CSA is not equally considered being 21 excluded from the decision-making processes (FAO, 2016b; Bruguere and Williams, 2017). Due to the lack 22 of monitoring programs, it is unknown how this tourism industry will respond to long-term changes driven 23 by climate change (Weatherdon et al., 2016). 24 25 12.5.2.5 Challenge and Opportunities 26 27 There is low evidence and high agreement that empower the local stakeholders (e.g., multilateral fisheries 28 agreements) improve the public awareness and simplify regulations and increase the flexibility and 29 sustainability of marine resources subjected to fisheries under future scenarios (Weatherdon et al., 2016; 30 Kalikoski et al., 2019). Ecosystem-Based Fishery Management (EBFM) arises as a suitable tool to minimize 31 the risk to climate change, avoid the degradation of the ecosystems and its services (Gullestad et al., 2017) 32 and maintain the long-term socioeconomic benefits when include climate complexity and the relationships 33 among species within the ecological systems (Long et al., 2015). There is high confidence that EbA is more 34 successful and feasible than hard coastal defences for the protection, management and restoration of ocean 35 and coastal ecosystems and their resources (Spalding et al., 2014; González and Holtmann-Ahumada, 2017; 36 Scarano, 2017). 37 38 There is high confidence that ecological and social resilience is improved by the presence of adequate 39 metrics evaluating adaptation measures that allow dynamic changes, increasing basic research and climate 40 data (Moreno et al., 2020b), the existence of early warning systems, improved local institutions, the 41 construction of adequate infrastructure, major funding for capacity building, and the enhanced engagement 42 and empowerment of women (FAO, 2016b; Harper et al., 2017; Frangoudes and Gerrard, 2018; Gallardo- 43 Fernández and Saunders, 2018; Leal Filho, 2018). 44 45 12.5.3 Water 46 47 CSA is one of the regions most affected by current and future hydrological risks to water security with an 48 increasing number of vulnerable people depending on water from mountain (high confidence) (Sections 4.3, 49 4.4, 4.5; Immerzeel et al., 2020; Viviroli et al., 2020; WWAP, 2020). Adaptation to changing water 50 availability is therefore a priority, but most efforts are documented only in the grey literature (e.g., 51 governmental documents, project reports) with highly variable standards of quality and evidence. Most of the 52 documented adaptation initiatives are in an early planning or implementation stage and evidence on 53 successful outcomes is quite limited (Berrang-Ford et al., 2021). However, the growing number of adaptation 54 initiatives across the CSA region has contributed to improved understanding of complex interlinkages of 55 climate change, human vulnerabilities, local policies, and feasible adaptation approaches (McDowell et al., 56 2019). 57 Do Not Cite, Quote or Distribute 12-58 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 12.5.3.1 Challenges and opportunities 2 3 In several regions of CSA, water scarcity is a serious challenge to local livelihoods and economic activities. 4 Particularly (seasonally) dry regions, partly with large populations and increasing water demand, exhibit 5 major water stress. These include the dry corridor in CA, coastal areas of Peru (SWS) and Northern Chile 6 (SWS), the Bolivian-Peruvian Altiplano (NWS, SAM), the Dry Andes of Central Chile (SWS),Western 7 Argentina and Chaco in Northwest Paraguay (SES), and Sertão in Northeast Brazil (NES) (high confidence) 8 (Kummu et al., 2016; Mekonnen and Hoekstra, 2016; Schoolmeester et al., 2018). In NWS and SWS, 9 downstream areas are increasingly affected by decreasing and unreliable river runoff due to rapid glacier 10 shrinkage (high confidence) (Table SM12.6; Carey et al., 2014; Drenkhan et al., 2015; Buytaert et al., 2017). 11 Many regions in CSA rely heavily on hydroelectric energy, and as a result of rising energy demand, 12 hydropower capacity is constantly extended (Schoolmeester et al., 2018). Worldwide, SA features the 13 second-fastest growth with about 5.2 GW additional annual capacity installed in 2019 (IHA, 2020). This 14 development requires additional water storage options, which entail the construction of large dams and 15 reservoirs with important social-ecological implications. River fragmentation and corresponding loss of 16 habitat connectivity due to dam constructions have been described for e.g., the NSA, SAM, NES and SES 17 (high confidence) (Grill et al., 2015; Anderson et al., 2018a) with important implications for freshwater 18 biota, such as fish migration (medium confidence) (Pelicice et al., 2015; Herrera-R et al., 2020). Furthermore, 19 examples in e.g., the NWS (Carey et al., 2012; Duarte-Abadía et al., 2015; Hommes and Boelens, 2018) and 20 SWS (Muñoz et al., 2019b) showcase unresolved water-related conflicts between local villagers, peasant 21 communities, hydropower operators and governmental institutions in a context of distrust and lack of water 22 governance (high confidence). 23 24 Increasing water scarcity is also shaped by poor water quality, which has barely been assessed in CSA. 25 Declining water quality can be observed e.g., due to intense agricultural and industrial activities in SWS, 26 SES and SSA (medium confidence) (Mekonnen et al., 2015; Gomez et al., 2021), mining in Andean 27 headwaters (NWS, SWS and Western SAM) and tropical lowlands (Eastern SAM and NSA) (medium 28 confidence) (Bebbington et al., 2015 risk and climate resilience; Vuille et al., 2018), urban domestic use 29 (Desbureaux and Rodella, 2019), decreasing meltwater contribution (Milner et al., 2017) and acid rock 30 drainages from recently exposed glacial sediments (Santofimia et al., 2017; Vuille et al., 2018). The level of 31 water pollution is often exacerbated by missing water treatment infrastructure and low governance levels 32 (medium confidence) (Mekonnen et al., 2015) with considerable negative implications for human health 33 (Lizarralde Oliver and Ribeiro, 2016). 34 35 Water scarcity risks are projected to affect a growing number of people in the near and mid-term future in 36 view of growing water demand in most regions (medium confidence: medium evidence, high agreement) 37 (Veldkamp et al., 2017; Schoolmeester et al., 2018; Viviroli et al., 2020), expected precipitation reductions 38 in Western and Northern SAM and SWS (medium confidence: medium evidence, medium agreement) 39 (Neukom et al., 2015; Schoolmeester et al., 2018), substantial vanishing of glacier extent in NWS, SAM and 40 SWS (Table SM12.6; Rabatel et al., 2018; Vuille et al., 2018; Cuesta et al., 2019; Drenkhan et al., 2019), and 41 increasing evaporation rates in CA (medium confidence) (CEPAL, 2017). Furthermore, flood risk is a serious 42 concern (Arnell et al., 2016) and expected to increase especially in NWS, SAM, SES and SWS in the mid 43 and long-term future (high confidence) (Arnell and Gosling, 2016; Alfieri et al., 2017). 44 Risks of water scarcity and flood are threatening people unevenly across the region. In CSA, about 26% (130 45 million people) of the population have no access to safe drinking water and strong disparities prevail 46 regarding its spatial distribution, e.g., in Chile 99% of the population have access, compared to 50% in Peru, 47 73% in Colombia, 52% in Nicaragua or 56% in Guatemala (high confidence) (UNICEF and WHO, 2019). 48 Inequalities can be further exacerbated by unregulated or privately owned water rights and allocation 49 systems (e.g., in Chile) (Muñoz et al., 2020a). The most vulnerable people belong to low-income groups in 50 rural areas and informal settlements of large urban areas (high confidence) (WWAP, 2020). 51 52 Considerable uncertainties remain concerning future hydrological risks that strongly depend on the 53 respective pathways of human intervention, management, adaptation and socioeconomic development. The 54 combination of (seasonally) reduced water supply, growing water demand, declining water quality, 55 ecosystem deterioration and habitat loss, and low water governance could lead to increasing competition and 56 conflict associated with high economic losses (high confidence) (Vergara et al., 2007; Vuille et al., 2018; 57 Desbureaux and Rodella, 2019). This situation threatens human water security on the long term and poses an Do Not Cite, Quote or Distribute 12-59 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 increasing risk to adaptation success in CSA (high confidence) (Drenkhan et al., 2015; Huggel et al., 2015b; 2 Urquiza and Billi, 2020a). 3 4 Important progress has been made on climate change and water management policies in combination with 5 more inclusive stakeholder processes. For instance, the implementation of NDCs in most countries of the 6 region provides an important baseline for improving water efficiency, quality and governance at multi- 7 sectoral level, and thus long-term adaptation planning (UNEP, 2015). 8 9 12.5.3.2 Main concepts and approaches 10 11 Adaptation in the water sector includes a broad set of responses to improve and transform, among others, 12 water infrastructure, ecosystem functions, institutions, capacity building and knowledge production, habits 13 and culture, and local-national policies (Section 4.6). 14 15 Most adaptive water management approaches in CSA centre around extending the water supply side 16 including large infrastructure projects. However, 'hard path' interventions are now strongly contested due to 17 negative effects exacerbating local water conflicts (Carey et al., 2012; Boelens et al., 2019; Drenkhan et al., 18 2019), potentially leading to increasing water demand, vulnerabilities and water shortage risks (Di 19 Baldassarre et al., 2018), and, hence, limiting adaptive capacity (high confidence) (Ochoa-Tocachi et al., 20 2019). More integrated approaches focus on multi-use of water storage with shared stakeholder vision, 21 responsibilities, rights and costs, as well as risks and benefits, and often integrating water and risk 22 management (Branche, 2017; Haeberli et al., 2017; Drenkhan et al., 2019). In this chapter, a feasibility 23 assessment was carried out for six major dimensions of multi-use water storage for the entire CSA (see Table 24 12.11). While geophysical and economic aspects allow for the implementation of water storage projects with 25 multi-use approach, the institutional, social and environmental dimensions pose a major barrier (see Section 26 12.5.3). Further demand-oriented approaches focus on incentives for the reduction of water use through 27 changes in people's habits, efficiency increase and smart water management (Gleick, 2002). These are 28 promoted in some regions, such as in CA and NWS (e.g., Colombia, Ecuador and Peru), to foster a 29 sustainable water culture (Bremer et al., 2016; Paerregaard et al., 2016). 30 31 Major attention has been put on nature-based solutions (NbS), i.e., catchment interventions that are inspired 32 and supported by nature and leverage natural processes and ecosystem services to contribute to the improved 33 management of water. NbS potentially enhances water infiltration, groundwater recharge and surface 34 storage, contributes to disaster risk reduction and can replace or complement grey (i.e., conventionally built) 35 infrastructure that is often socio-environmentally contested (WWAP, 2018). Some examples include the 36 reactivation of ancestral infiltration enhancement systems in the Peruvian Andes (NWS) (Ochoa-Tocachi et 37 al., 2019), the use of erosion control structures in the Bolivian Altiplano (SAM) (Hartman et al., 2016), and 38 the potential improvement of drinking water quality and flood risk reduction in urban areas of CSA (Tellman 39 et al., 2018, Section 12.5.5.3.2). Additionally, NbS in combination with ecosystem and community-based 40 adaptation potentially generate important co-benefits including increasing water security and the attenuation 41 of social conflicts in Chile (SWS) (Reid et al., 2018), water conservation in coastal Peru (NWS), and flood 42 protection in Guyana (NSA) (medium confidence: medium evidence, medium agreement) (Spencer et al., 43 2017). However, evaluation of implementation success of NbS is often hampered by limited evidence on 44 actual benefits (WWAP, 2018). 45 46 In recent years, the inclusion of Indigenous knowledge (IK) and local knowledge (LK) into current 47 adaptation baselines has gained increasing attention, particularly in regions with a high share of Indigenous 48 Peoples (NWS, SAN, SWS, NSA) (high confidence) (Reyes-García et al., 2016; Schoolmeester et al., 2018; 49 McDowell et al., 2019). One example is the adapted use of agrobiodiversity when dealing with more 50 frequent and intense tidal floods in the Amazon delta (NSA) (Vogt et al., 2016). In another context, IK and 51 LK have been considered for the evaluation of water scarcity and glacier lake outburst flood risks in Peru 52 (NWS) (Motschmann et al., 2020b). Additionally, local citizen science based initiatives (Buytaert et al., 53 2014; Tellman et al., 2016; Njue et al., 2019) can support the production of multiple knowledge with flexible 54 and extensive data collection. Important questions centre around how to integrate IK, LK and other types of 55 knowledge from the early planning stages on, to achieve enhanced or transformational adaptation building 56 on co-produced knowledge (Kates et al., 2012; Klenk et al., 2017). NbS combined with community Do Not Cite, Quote or Distribute 12-60 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 engagement and integration of diverse knowledge can foster transformational adaptation of social-ecological 2 systems (Palomo et al., 2021). 3 4 5 Figure 12.13: Overview map of observed glacier changes, associated impacts, adaptation and policy efforts across the 6 Andes. (a) Selected impacts from glacier shrinkage. (b) Selected adaptation efforts (see upper-right map for the location 7 of each adaptation measure), (c) Policies and glacier inventory: NDC = submission year(s) of Nationally Determined 8 Contributions (u = update), CCL = climate change law, GLL = glacier law (i = initialized framework), INV = last 9 national glacier inventory. The explicit mention of glaciers, snow and mountain ecosystems within each law/inventory 10 is highlighted with the corresponding symbols (grey colour = not come into force). (d) Glacier area (km²) according to 11 last national inventory. (e) Glacier area change (%/year) according to the baseline of the last national inventory. (f) 12 Geodetic glacier mass balance (m w.e./year) and error estimate (±m w.e./year) retrieved from Dussaillant et al. (2019). 13 nd = no data available. Further details can be found in the Appendix in Table SM12.6. 14 15 16 12.5.3.3 Policies, governance and financing 17 18 National policies on climate change, water protection, regulation and management laws are important focal 19 areas of adaptation in the water sector (Section 4.7). Notable in the jurisdiction field is the Glacier Protection 20 Law in place in Argentina (2010-2019), and under construction in Chile (since 2005). This first glacier law 21 in the world represents a milestone for high-mountain conservation but is also criticized for hindering 22 effective disaster risk adaptation measures and excluding local socioeconomic needs (Anacona et al., 2018). 23 Furthermore, the first Framework Law on Climate Change was implemented in Peru (2018), and is 24 underway in Colombia, Chile and Venezuela (Figure 12.13; Table SM12.6). Overarching regional Do Not Cite, Quote or Distribute 12-61 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 institutions (e.g., OAS (2016)) and most countries in CSA promote a move towards more integrative and 2 sustainable management of water resources through new legislations and financing mechanisms. For 3 instance, new water laws including principles of Integrated Water Resources Management (IWRM) have 4 entered into force, e.g., in Nicaragua (2007), Peru (2009), Ecuador (2014) and Costa Rica (2014) or are 5 underway, such as in Colombia (since 2009). However, current realities in all regions show major challenges 6 in implementing IWRM mechanisms and policies, related but not limited to political and institutional 7 instabilities, governance structures, fragmented service provision, lack of economies of scale and scope, 8 corruption and social conflicts (high confidence) (WWAP, 2020). 9 10 Many water-related conflicts in CSA are rooted in inequitable water governance that excludes water users 11 from decisions on water allocation (high confidence) (Drenkhan et al., 2015; Vuille et al., 2018). In turn, 12 inclusive water regimes leverage long-term adaptation planning. These have been addressed in some national 13 strategies, such as in Brazil (Ministry of Environment of Brazil, 2016a). At the local level, a decentralized 14 and participatory bottom-up water governance model was induced by civil society and research institutions 15 to foster rainwater harvesting technologies reducing drought risk in semi-arid Brazil (NES) (Lindoso et al., 16 2018). 17 18 Water fund programs can generate important co-benefits for Sustainable Development contributing to 19 improved governance and conservation of watershed systems in CSA. Nevertheless, only a few experiences 20 have been evaluated as successful due to insufficient implementation, low decision-making of some 21 stakeholder groups and poor evidence-based approaches (medium confidence) (Bremer et al., 2016; Leisher 22 et al., 2019). Furthermore, financing mechanisms that produce incentives for sustainable water management 23 have been promoted, tested or implemented. Payments for Ecosystem Services (PES) for water provision 24 represent such an example and have been implemented across CSA since the 1990s (Grima et al., 2016). 25 26 Only about 50­70% of required financial resources are currently allocated per year to meet the national 27 targets in the water, sanitation and hygiene (WASH) sector for the Sustainable Development Agenda (SDG 28 6) in several regions of CSA. This share drops down to less than 50% in NSA (Venezuela) and SES 29 (Argentina, Uruguay, Paraguay), except for Panama in CA allocating more than 75% of required financial 30 resources. For the implementation of NbS, evidence suggests that the overall expenditure remains well below 31 1% of total investment in water resources management infrastructure (WWAP, 2018). These funding deficits 32 pose important limitations for future water provision, adaptation to changing water resources, and the 33 achievement of the SDGs by 2030 (high confidence) (WHO, 2017). 34 35 12.5.3.4 Successful adaptation and limitations 36 37 Although a growing body of adaptation initiatives exists for CSA, evidence on effectiveness is scarce. In 38 many parts of CSA the level of success of respective adaptation measures depends much on the governance 39 of projects and stakeholder-based processes and is closely related to their effectiveness, efficiency, social 40 equity and socio-political legitimacy (high confidence) (Adger et al., 2005; Rasmussen, 2016b; Moulton et 41 al., 2021). Several Payments for Ecosystem Services experiences across CSA have been described as 42 successful measures for watershed conservation and adaptation (high confidence). An example of success 43 represents the Quito water fund in Ecuador which aims at improving the city's water quality by integrating 44 public and private stakeholder interests with ecosystem conservation and local community development 45 since the 2000's (Bremer et al., 2016; Grima et al., 2016) (case study 12.6.1). At the same time, in 46 Moyobamba in Peru the development of a watershed protection program was leveraged by a multi- 47 stakeholder platform process that enabled deep social learning (Lindsay, 2018). In turn, initiatives that do not 48 consider the entire set of social-ecological dimensions and dynamics of adaptation or unintentionally 49 increase vulnerabilities of human or natural systems, are at risk to lead to reduced outcomes (McDowell et 50 al., 2021) or maladaptation (Reid et al., 2018; McDowell et al., 2019; Eriksen et al., 2021). However, 51 systematic assessments of maladaptation in the water sector have barely been provided for CSA. 52 53 In CSA, only limited information on limits of adaptation in relation to water is available, for instance on 54 possible path dependency of institutions and associated resistance to change (Barnett et al., 2015). Examples 55 of soft adaptation limits (i.e., options to avoid intolerable risks currently not available) include the lack of 56 trust and stakeholder flexibility, associated with unequal power relations that lead to reduced social learning, 57 and poor outcomes for improved water management, as reported in e.g., NWS (Lindsay, 2018). An example Do Not Cite, Quote or Distribute 12-62 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 for hard adaptation limits (i.e., intolerable risks cannot be avoided) in the region is the loss of livelihoods and 2 cultural values associated with glacier shrinkage in NWS (Jurt et al., 2015). 3 4 Most barriers to advance adaptation in CSA correspond to soft limits associated with missing links of 5 science-society-policy processes, institutional fragilities, pronounced hierarchies, unequal power relations 6 and top-down water governance regimes (high confidence). One example is the abandonment of hydrological 7 long-term monitoring sites within tropical Andean ecosystems (paramo) in Venezuela (Rodríguez-Morales et 8 al., 2019) due to the lack of governmental support within a political crisis. In that regard, the collection and 9 availability of consistent hydroclimatic and socioeconomic data at adequate scales represent an important 10 challenge in CSA. Major adaptation barriers are furthermore reported from Central Chile in the context of a 11 mega-drought since 2010, related to socioeconomic characteristics and a deficient bottom-up approach to 12 public policy informing and development (Aldunce et al., 2017). These gaps could be bridged by 13 strengthening transdisciplinary approaches at the science-policy interface (Lillo-Ortega et al., 2019) with 14 blended bottom-up and top-down adaptation to include scientific knowledge with impact and scenario 15 assessments into local adaptation agendas (Huggel et al., 2015b). For instance, a new allocation rule for the 16 Laja reservoir in Southern Chile (SWS), based on consistent water balance modelling results, could inform 17 policy and water management and potentially improve local water management and reduce water conflicts 18 on the long term (Muñoz et al., 2019b). 19 20 12.5.4 Food, Fibre and other Ecosystem Products 21 22 The CSA region globally has the greatest agricultural land and water availability per capita. With 15% of the 23 world's land area, it receives 29% of global precipitation and has 33% of globally available renewable 24 resources (Flachsbarth et al., 2015). Agricultural commodities (coffee, bananas, sugar, soybean, corn, 25 sugarcane, beef livestock) are some of the highest users of ecosystem resources such as land, water, nutrients 26 and technology. These exports have gained importance in the past two decades as international trade and 27 globalization of markets have shaped the global agri-food system. However continuous overuse on the 28 environment might account for resource depletion (deforestation, land degradation, nutrient depletion, 29 pollution), affecting the natural capital base. The effects of climate change on humans, via ecological 30 systems, exacerbate the impact related to depletion of ecosystem services (Scholes, 2016; IPBES, 2018b; 31 Castaneda Sanchez et al., 2019; Clerici et al., 2019; Tellman et al., 2020; Pacheco et al., 2021). 32 33 12.5.4.1 Challenges and opportunities 34 35 Even though there are large improvements in food availability in several regions, there is also a tendency of 36 a decline in food self-sufficiency in many countries (Porkka et al., 2013; Rolando et al., 2017). Drought 37 conditions in Central America and the Caribbean increased in line with climate model predictions (Herrera et 38 al., 2018a). The direct social and economic consequences for the sector are evident in Central America's so- 39 called Dry Corridor with a growing dependence on food imports (Porkka et al., 2013) and these degrees of 40 dependency make the region more vulnerable to price variability, climatic conditions (Bren d'Amour et al., 41 2016; ECLAC, 2018) and therefore, to food insecurity if adaptation actions are not taken (high confidence) 42 (Porkka et al., 2013; Bren d'Amour et al., 2016; López Feldman and Hernández Cortés, 2016; Eitzinger et 43 al., 2017; Imbach et al., 2017; Lachaud et al., 2017; Harvey et al., 2018; Niles and Salerno, 2018; del Pozo et 44 al., 2019; Alpízar et al., 2020; Anaya et al., 2020). 45 46 Given these circumstances, some regions in CSA (Andes region and Central America) will just meet, or fall 47 below, the critical food supply/demand ratio for their population (Bacon et al., 2014; Barbier and Hochard, 48 2018b). Meanwhile, the more temperate part of South America in the south is projected to have agricultural 49 production surplus (low confidence) (Webb et al., 2016; Prager et al., 2020). The challenge for this region 50 will be to retain the ability to feed and adequately nourish its internal population as well as making an 51 important contribution to the food supplies available to the rest of the world. 52 53 The access of agricultural products from the region to other markets might be conditioned on the adoption of 54 low-carbon agriculture measures. Achieving net-zero emissions while improving standards of living is 55 possible but requires developing transition policy frameworks to attain the target (Frank et al., 2019; 56 Mahlknecht et al., 2020; Cárdenas et al., 2021). 57 Do Not Cite, Quote or Distribute 12-63 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 12.5.4.2 Governance and barriers for adaptation 2 3 The governance of adaptation for CSA implies modifying agricultural, socio-economic and institutional 4 systems in response to and in preparation for actual or expected impacts of climate variability and change, to 5 reduce harmful effects and exploit beneficial opportunities (high confidence). CSA agriculture has a diversity 6 of systems and segments of producers. While small-scale farmers have a big contribution to food production 7 and food security, especially in developing economies, they face global policies oriented towards global 8 commodity markets (Knapp, 2017; Fernández et al., 2019). Climate action initiatives that consider CSA's 9 high levels of poverty and inequality to reduce these pervasive problems are central for adapting the region 10 (Crumpler et al., 2020; Locatelli et al., 2020). 11 12 Since AR5, important advances at institutional level are observed based on the development and 13 implementation of national adaptation plans for the agriculture and forestry sector among countries. 14 Adapting to climate change entails the interaction of decision-makers, stakeholders, and institutions at 15 different scales of government from the local to the national. The Climate-Adapted Sustainable Agriculture 16 Strategy for the region of the Central American Integration System (EASAC) of the Central American 17 Agricultural Council of Ministers of Agriculture, constitutes a valuable example of how undertake climate 18 action in the agricultural sector, as a block of countries and in an intersectoral manner, to enhance results and 19 make better use of resources (IICA, 2019). 20 21 In Brazil, the Low Carbon Agriculture program (Programa ABC) funds practices for reducing GHG 22 emission in the sector (Government of Brazil, 2012), allocating about 15% of the total agriculture official 23 finance portfolio, although it faces challenges to advance (Souza Piao et al., 2021). Costa Rica offers an 24 example on how reforestation can help achieve Paris Agreement objectives. Reforestation through natural 25 regeneration on abandoned pastures boosted forest cover from 48% in 2005 to 53.4% in 2010 (Reid et al., 26 2019; Cárdenas et al., 2021). Some key success factors included a strong institutional context, fiscal and 27 financial incentives for reforestation, conservation measures such as payment for environmental services, 28 cattle ranch subsidy reform, and a historically strong enforcement and focus on land titles that favoured the 29 restoration of lands. Uruguay offers another example, with the farm sector contribution of 32.8% of all 30 exports and 73.8% of the country's emissions, so decarbonisation is not just an environmental issue but an 31 economic competitiveness one as well. In the INDCs submitted to the UNFCCC in 2015, Uruguay set a 32 specific target for the agriculture sector to reduce enteric methane emissions intensity per kilogram of beef 33 (live-weight) by 33% to 46% in 2030 through improving efficiency of beef production by controlling the 34 grazing intensity to increase animal intake, reproductive efficiency, and daily weight gain (Picasso et al., 35 2014). 36 37 It is relevant to generate conditions for the development of sustainable agricultural practices in a frame 38 where factors associated with climate have become important for producers, given recent experiences of 39 drought and lack of water (high confidence) (Clarvis and Allan, 2014; Roco et al., 2016; Hurlbert and Gupta, 40 2017; Pérez-Escamilla et al., 2017; Cruz et al., 2018; Zúñiga et al., 2021). Solutions that consider relevant 41 drivers that have demonstrated positive effect in diffusion of adaptation strategies are more efficient (Table 42 12.8). Some conditions such as the promotion of education programs; participation in cooperatives; credit 43 access; land tenure security can help in this task. In the same line, in CSA some elements such as technology 44 and information access, and local knowledge, reinforce climate change adaptation (Khatri-Chhetri et al., 45 2019; Piggott-McKellar et al., 2019). As is stated in Table 12.8 barriers of different origin persist for climate 46 change adaptation in the region increasing vulnerability of farming systems and rural livelihoods. 47 48 Limited information regarding cost-benefit analyses of adaptation is available in the region as well as 49 avoiding maladaptation effects and promoting site-specific and dynamic adaptation options considering 50 available technologies (medium confidence) (Roco et al., 2017; Zavaleta et al., 2018; Ponce, 2020; Shapiro- 51 Garza et al., 2020). 52 53 Climate Information Services has an important role in climate change adaptation and there is a recognized 54 gap between climate science and farmers (high confidence) (Vaughan et al., 2017; Loboguerrero et al., 2018; 55 Tall et al., 2018; Thornton et al., 2018; Ewbank et al., 2019). Such services should address the challenges of 56 ensuring that climate information and advisory services are relevant to the decisions of small-holder and 57 family farmers, providing timely climate services access to remote rural communities with marginal Do Not Cite, Quote or Distribute 12-64 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 infrastructure and ensuring that farmers own climate services and shape their design and delivery. An 2 interesting case facing this gap is the implementation of local technical agro-climatic committees in 3 Colombia which allow to share and to validate climatic and weather forecasts; and crop model results to 4 seasonal drought events (Loboguerrero et al., 2018). Another example is the web service, AdaptaBrasil- 5 MCTI, forecasting the risk of climate change impact on strategic sectors (e.g., food, energy, water) in Brazil 6 (Government of Brazil and Ministry of Science Technology and Innovation Secretariat of Policies and 7 Programs, 2021). 8 9 Barriers to financial access are present in the region restricting effective adaptation to extreme weather 10 events (high confidence) (Chen et al., 2018; Fisher et al., 2019; Piggott-McKellar et al., 2019; Vidal Merino 11 et al., 2019; de Souza Filho et al., 2021). In 2014, the penetration rate of this type of insurance in the region 12 averaged 0.03% of GDP, and a few countries dominate the market (Brazil, Argentina). Beyond these three 13 countries, some initiatives also exist in Uruguay, Paraguay, Chile and Ecuador. In most Latin American and 14 Caribbean countries, the public sector plays an important role in providing insurance or reinsurance and 15 coexists with private sector companies (Cárdenas et al., 2021). Insurance protections represent a strategy to 16 transfer climate risk to protect the wellbeing of vulnerable small farmers and accelerate uptake (recovery) 17 after a climate-related extreme weather event. Lack of finance and proper infrastructure is compounded by 18 limited knowledge of sustainable farming practices and high rates of financial illiteracy (high confidence) 19 (Hurlbert and Gupta, 2017; Piggott-McKellar et al., 2019). 20 21 Insufficient access to digital services and technologies further widens the gap between the rural poor and 22 more urban populations of Latin America and the Caribbean (medium confidence: insufficient evidence, high 23 agreement). In turn, these factors compromise productivity and competitiveness. Support for this group can 24 be focused on both economic competitiveness and social development. Finally, to align identified adaptation 25 options as a priority for achieving future food security in the NDCs of CSA countries to mitigation 26 commitments, it will be essential to highlight synergies by generating evidence (national research) in relation 27 to progress towards increasing productivity, resilience and reducing GHG; and also demonstrating its added 28 value as a development initiative (Rudel et al., 2015 sustainable; Loboguerrero et al., 2019). 29 30 12.5.4.3 Adaptation options 31 32 In order to contextualize the adaptation options at the regional level, the majority of the NDC of the CSA 33 countries reported the observed and/or projected climate-related hazards: occurrence of droughts and floods 34 (80% of countries each), followed by storms (45%) and landslides (30%), as well as extreme heat, wildfire 35 and invasion by pests and non-native species in agriculture (25% each) (Crumpler et al., 2020). 36 37 Main adaptation options for climate change in the region include preventive measures against soil erosion; 38 climate-smart agriculture which provide a framework for synergies between adaptation, mitigation and 39 improved food security; climate information systems; land use planning; shifting plantations in high altitude 40 to avoid temperature increases and plagues; improved varieties of pastures and cattle (Lee et al., 2014; Jat et 41 al., 2016; Crumpler et al., 2020; Moreno et al., 2020a; Aragón et al., 2021). Agricultural technologies are not 42 necessarily changing, but the economic activity is shifting to accommodate increasing climate variation and 43 adapt to changes in water availability and ideal growing conditions (high confidence) as is observed in 44 Argentina, Colombia and Brazil (McMartin et al., 2018; Rolla et al., 2018; Sloat et al., 2020; Gori Maia et 45 al., 2021). Coffee plantations are moving further up mountain regions with the land at lower elevations 46 converted for other uses. In Brazil, crop modelling suggests the need for the development of new cultivars, 47 with a longer crop cycle and with higher tolerance to high temperatures, a necessary technological advance 48 for maize, an essential staple crop, to be produced in the future. Additionally, irrigation becomes essential for 49 sustaining productivity in adverse climate change scenarios in several regions of CSA (McMartin et al., 50 2018; Lyons, 2019; Reay, 2019). 51 52 Livestock production is for small farmers one of the main sources of protein and contributes to food security 53 (Rodríguez et al., 2016). The importance of this sub-sector in CSA, will continue to increase as the demand 54 for meat products does as well in the coming years, driven by growing incomes in the region (OECD and 55 FAO, 2019). However, the increase in animal production has been associated with land degradation, 56 triggered by the conversion of native vegetation to pastureland and aggravated by overgrazing and 57 abandoning of the degraded pastures (Baumann et al., 2017; ECLAC, 2018; Müller-Hansen et al., 2019). Sá Do Not Cite, Quote or Distribute 12-65 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 et al. (2017) simulated the adoption of agricultural systems based on Low-Carbon Agriculture (LCA) 2 strategies towards 2050. According to the simulation, the adoption of LCA strategies in the SA region can 3 alter the growing trend of Land Use and Land Use Change emissions and at the same time, it can increase 4 meat production by 55Mt for the entire period (2016­2050). The restoration of degraded pasture and 5 livestock intensification account for 71.2%, and integrated crop-livestock-forestry system contributes 28.8% 6 of total meat production for the entire period. These results indicate that combined actions in agricultural 7 management systems in SA, can result in synergistic responses that can be used to make agriculture and 8 livestock production an important part of the solution of global climate change and advance food security 9 (medium confidence: insufficient evidence and high agreement) (Zu Ermgassen et al., 2018; Pompeu et al., 10 2021). Crop-Livestock-Forestry-Systems are also important for climate change adaptation as they provide 11 multiple benefits, including the coproduction of food, animal feed, organic fertilizers and soil organic carbon 12 sequestration (Sharma et al., 2016; Rodríguez et al., 2021), achieving mitigation and adaptation goals (high 13 confidence) (Picasso et al., 2014; Modernel et al., 2016; Modernel et al., 2019; Rolla et al., 2019; Locatelli et 14 al., 2020). A recent analysis of agroforestry in Brazil, has shown positive and relevant impacts on the 15 heads/pasture area rate in livestock production and that the system may have also stimulated a shift toward 16 other production activities with higher gross added value (Gori Maia et al., 2021). Agroforestry has also 17 proven to have protective benefits to obtain more stable, less fluctuating yields due to climate damages in 18 coffee production (high confidence) (Bacon et al., 2017; Durand-Bessart et al., 2020; Ovalle-Rivera et al., 19 2020). In the same way, the production of plant-based fibre can be less vulnerable to economic and climatic 20 variability through farming systems diversification. Textile fibre crops for the case of cotton include crop 21 rotation, agroecological intercropping and agroforestry (Oliveira Duarte et al., 2019). 22 23 Adaptation strategies also concern Indigenous agriculture, i.e., the vast majority of the 44 million 24 Amerindians (CEPAL, 2014). Indigenous knowledge and local knowledge (IK and LK) can play an 25 important role in adaptation (Zavaleta et al., 2018). On one hand, they preserve the conservation of a very 26 rich agrobiodiversity that is likely to meet the challenges of climate change (high confidence) (Carneiro da 27 Cunha and Morim de Lima, 2017; Magni, 2017; Emperaire, 2018; Donatti et al., 2019) and on the other 28 hand, the sustainability of large territories that assure their livelihood (Singh and Singh, 2017; Mustonen et 29 al., 2021). In the Andes, ancient technologies increased the quantity of crops produced and allowed for 30 coping with climatic changes and water scarcity, while nutrition conditions were improved (high confidence) 31 (López Feldman and Hernández Cortés, 2016; Parraguez-Vergara et al., 2018; Carrasco-Torrontegui et al., 32 2020 food). Also, fire prevention management, protection against forest and biodiversity loss, are recognized 33 as important elements in Indigenous knowledge (Mistry et al., 2016; Bowman et al., 2021). 34 35 36 Table 12.8: Recent studies related to climate change adaptation of agricultural systems and its determinants in the CSA 37 Region. Authors, Countries Sampl Approach Crop Adaptation Main Main Main year e size of the systems strategies drivers barriers barriers (n) study promoting limiting detected climate climate change change adaptation adaptation de Souza 175 Quant. Cattle Integrated Credit Lack of Lack of Filho et Brazil farmers crop-livestock access resources agricultur al. (2021) and livestock- Extension al market forestry services access systems strategies Magalhãe Several Farm Previous Inadequate Infrastruc crops management experience infrastructu ture s et al. Brazil 94 Qual. with risks re limiting Low opportuni (2021) purchasing ties power Carrer et Brazil 175 Quant. Several Agricultural Schooling Higher risk Limited al. (2020) crops insurance propensity financial Do Not Cite, Quote or Distribute 12-66 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report Technical market assistance access Quiroga Nicaragua 212 Quant. Coffee Several Farm size Limited Absence et al. Quant. adaptation Awareness access to of (2020) Qual. measures of climate rain-water climate Quant. change change Quant. Schooling education Bro et al. Nicaragua 236 Quant. Coffee Crop Schooling Household Institutio (2019) Soil and Participation size nal Quant. water in framewor Quant. cooperatives k to Radio promote cooperati ves Leroy Venezuela Several Irrigation Perception Degradatio Ineffectiv (2019) crops in management of water n of fragile eness of and 73 high scarcity areas local altitudes Local institutio Colombia knowledge ns Cherubin Several Agroforestry Improving Degradatio Lack of crops systems soil quality n of crop et al. Colombia 6 and and biota convention diversific pasture al pasture ation (2019) Costa Coffee, Affordabilit Lack of beans y of adaptatio Harvey et Rica, 860 and Several Awareness adaptation n al. (2018) Honduras maize adaptation of climate practices involving and practices change agroecolo gical and Guatemala socioeco nomic contexts Chen et Costa Rica 559 Several Intensificatio Access to Land Lack of al. (2018) and crops n and weather renting crop and Nicaragua diversificatio information practices n Participation diversific in ation organization s Credit access Farming experience Vidal Several Water Farm size Limited Lack of crops management Capital access to site- Merino et Peru 137 Irrigated off-farm specific proportion activities design of al. (2019) Small interventi cultivated ons area Meldrum Potato, Diversificatio Weather Loss to Lack of quinoa traditional resilience et al. Bolivia 193 and n of crop information knowledge and others actions to (2018) portfolio Do Not Cite, Quote or Distribute 12-67 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report expand and maintain variety portfolio Lan et al. Nicaragua 180 Quant. Cocoa Crop Schooling Lack of Income (2018) management Household income inequalit size y Farm size Gaps of profitabil ity of practices Benefits of practices depends of its costs Kongsag 125 Qual. Maize Alley Schooling Land tenure Lack of er (2017) Belize cropping Market land distance tenure Degradatio Lack of n of fragile market areas access Lack of trust Schembe 5485* Quant. Several Agroforestry Financing High Adaptatio rgue et Brazil crops systems Presence of potential n al. (2017) associations for condition Credit agriculture ed by access Lack of agricultur climate al, information socioeco nomic and climatic condition s Guatemala Coffee Ecosystem Schooling Lack of and based Age access to Harvey et , Honduras 300 Quant. maize adaptation Farming Lack of training al. (2017) and Costa experience land tenure and Access to finance Rica technologica l support Roco et Chile 665 Quant. Several Water Farm size Locations Lack of al. (2016) crops management Access to Age availabili weather ty and information access to climate change informati on Mussetta Argentina 41 Qual. Vine Crop and Organization Water Lack of and and water of producers allocation water others management system managem Do Not Cite, Quote or Distribute 12-68 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report Barriento Labour ent and s (2015) availability distributi Knowledge on and strategies information access Technology access 1 Table Notes: 2 *: municipalities; Quant.: mainly quantitative; Qual.: mainly qualitative. 3 4 5 12.5.5 Cities, Settlements and Infrastructure 6 7 CSA is the second most urbanized region of the world, with 5 megacities and half of urban population in 129 8 secondary cities (UNDESA, 2019), huge metropolitan areas concentrated on the coast and an increasing 9 number of small cities by the sea (Barragán and de Andrés, 2016). Besides the many climatic events 10 threatening urban areas in the region (extreme heat, droughts, heavy storms, floods, landslides), cities by the 11 coast are also exposed to sea level rise (SLR) (Section 12.3; Figure 12.6; Dawson et al., 2018; Leal Filho et 12 al., 2018; Le, 2020). Main determinants of urban vulnerability assessed in the region are poor and unevenly 13 distributed infrastructure, housing deficits, poverty, informality and the occupation of risk areas, including 14 low elevation coastal zones (Section 12.3). Those features of urban systems increase the risks to health, 15 ecosystems and its services, water, food and energy supply (Section 12.4). Impacts of climate events on 16 urban water supply, drainage and sewer infrastructures are the most reported in the region (Section 12.3; 17 Figure 12.9). 18 19 12.5.5.1 Challenges and opportunities 20 21 Inequality, poverty and informality shaping cities in the region increase vulnerability to climate change (high 22 confidence) (Romero-Lankao et al., 2014; Rasch, 2017; Filho et al., 2019), and can hinder adaptation 23 (Section 12.5.7.1), while interventions addressing these social challenges and the existing development 24 deficits (e.g., build or improve infrastructure and housing applying climate-adapted patterns), can go hand in 25 hand with adaptation and mitigation (medium confidence: high agreement, medium evidence) (Section 26 12.5.7.3; Creutzig et al., 2016; Le, 2020; Satterthwaite et al., 2020). Over 20% of urban population in LAC 27 lives in slums and many in other forms of precarious and segregated neighbourhoods, settled in risk areas 28 and lacking infrastructure (Rasch, 2017; UN-Habitat, 2018; Rojas, 2019). This vulnerable condition is 29 boosted by unstable political and governmental institutions, which recurrently suffer from corruption, weak 30 governance and reduced capacity to finance adaptation (Rasch, 2016). Facing governance challenges by 31 including diverse stakeholders and encouraging and learning from community-based experiences has been 32 also an opportunity to improve adaptation strategies (Archer et al., 2014). The Regional Climate Change 33 Adaptation Plan of Santiago is an example of this (Krellenberg and Katrin, 2014). 34 35 12.5.5.2 Governance and Financing 36 37 Lack of a high multilevel and intersectoral governance capacity with strong multi-players horizontal and 38 vertical coordination and long-term support are limiting adaptation in the region (high confidence) 39 (Anguelovski et al., 2014; Bai et al., 2016; Chu et al., 2016; Schaller et al., 2016; Miranda Sara et al., 2017). 40 The ability to enrol stakeholders and include community based initiatives can be determinant for adaptation 41 success particularly considering its impact in the decision-making arena (high confidence) (Section 12.5.8.1; 42 Section 6.4; Anguelovski et al., 2014; Archer et al., 2014; Chu et al., 2017; Rosenzweig et al., 2018) . 43 44 Lima's Climate Action Strategy is an example (Metropolitan Municipality of Lima, 2014). It was approved 45 after a participatory and consultative process with the technical group on climate change from the 46 Metropolitan Environmental Commission, focusing on the reduction of water vulnerabilities to drought and 47 heavy rain, on the basis of which 10 (out of 51 with Callao) Lima districts municipalities are developing and 48 starting to implement their adaptation measures (Foro Ciudades Para la Vida, 2021). In 2021 Lima 49 Municipality also approved its Local Climate Change Plan (Metropolitan Municipality of Lima, 2021) under 50 a similar process. The engagement of local players was central to spreading and mobilizing different types of Do Not Cite, Quote or Distribute 12-69 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 knowledge and creating networks able to support adaptation (Section 12.6.3; Miranda Sara and Baud, 2014; 2 Miranda Sara et al., 2017) . The inclusive process is also a goal on the example of Chile Municipalities 3 Network Facing Climate Change (RedMuniCC) engaged in developing participatory strategic plans for 4 climate adaptation and mitigation (RedMuniCC, 2021). 5 6 New forms of financing and leadership focused on community-based approaches have been developed to 7 overcome the funding challenge and enable adaptation in the region (medium confidence: medium evidence, 8 medium agreement) (Castán Broto and Bulkeley, 2013; Archer et al., 2014; Paterson and Charles, 2019). 9 Also systems for measuring, reporting and verifying adaptation financing, as in Colombia (Guzmán et al., 10 2018), as much as national legislation geared to adaptation can help access funds. Peruvian Law on the 11 Retribution Mechanism of Eco-Systemic Services and Code (Miranda Sara and Baud, 2014; MINAM Peru, 12 2016) in addition to the Ley Marco de la Gestión y Prestación de los Servicios de Saneamiento and its Code 13 (Ministerio de Vivienda, 2017), allowed the potable water companies to add 1% to the tariff to guarantee 14 ecosystem services, water treatment and reuse with green infrastructure. Another 4% of tariffs go to develop 15 and implement adaptation plans and measures (Government of Peru, 2016). 16 17 12.5.5.3 Adaptation options in urban design and planning 18 19 Both the shape and activities of the city have an impact on carbon emissions, adaptation and mitigation 20 opportunities (high confidence) (Raven et al., 2018; Satterthwaite et al., 2018). Combining urgent measures, 21 strategic action (Chu et al., 2017) to long-term planning is central for a transformative adaptation and to 22 avoid maladaptation (Filho et al., 2019). Urban planning, considering climate risk assessments, and 23 regulation (e.g., land-use and building codes), including climate-adapted parameters, are central to 24 coordinate and foster private and public investments in adaptation, reducing risks related to the built 25 environment conditions (infrastructure and buildings) and the occupation of risk areas (e.g., threatened by 26 floods and landslides) (Rosenzweig et al., 2018). Lack of information at local scale, human resources and 27 clear liability for climate change response planning can limit adaptation (Aylett, 2015). 28 29 Strategic adaptation approaches have been adopted by many cities in dealing with the multilevel and 30 intersectoral complexity of urban systems, with gains in fostering leadership and facing the predominant 31 pattern of uneven urban development in the region (medium confidence: limited evidence, high agreement) 32 (Chu et al., 2017). Medellin's metropolitan green belt, for example, focuses on problems such as irregular 33 settlements, inequality and poor governance, articulating programs and projects of the Municipality of 34 Medellin and the municipalities of the Vale do Aburra in a strategic long-term planning. Places with 35 informal and precarious settlements were aimed to be transformed with the belt's integration areas: eco parks 36 and eco-gardens (Alcaldía de Medellín, 2012; Chu et al., 2017). 37 38 12.5.5.3.1 Housing, informality and risk areas 39 Informality and precariousness in housing is one of the most sensitive issues for adaptation in CSA cities 40 (medium confidence: medium evidence, high agreement) (Satterthwaite et al., 2018; UN-Habitat, 2018). 41 Housing deficit in 2009, as a regional baseline, estimated that 37% of households suffered from quantitative 42 or qualitative deficiencies, due to the high cost of housing and the incidence of poverty (Blanco Blanco et al., 43 2014; McTarnaghan et al., 2016; NU CEPAL et al., 2016; Vargas et al., 2018a; Rojas, 2019). 44 45 Policies and programs have been implemented accumulating good practices and reducing the percentage of 46 population in informal and precarious settlements (33.7% in 1990 to 21% in 2014) (NU CEPAL et al., 2016; 47 Satterthwaite et al., 2018; Teferi and Newman, 2018; UN-Habitat, 2018). Slum Upgrading and built- 48 environment interventions (housing and infrastructure improvement and provision) in informal settlements 49 can enhance adaptation (high confidence) (Teferi and Newman, 2018; Núñez Collado and Wang, 2020; 50 Satterthwaite et al., 2020) while reducing floods, landslides and cascading impacts of storms, floods and 51 epidemics, as observed on the "incremental housing approach" in Quinta Monroy (Rojas, 2019) and the 52 "social urbanism" in Medellin (Garcia Ferrari et al., 2018). 53 54 The climate adaptation plans of several large CSA cities include efficient land use and occupation planning 55 and urban control systems (comprising regulation, monitoring), fostering interlocution with housing and 56 environmental policy (by means of intersectoral and multilevel governance), inhibiting and reducing the 57 occupation of risk areas (mainly flooding and landslides risks); increasing population density in areas already Do Not Cite, Quote or Distribute 12-70 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 served by infrastructure; expanding slums urbanization and technical assistance programs for improvements 2 and expansion of social housing (high confidence) (Municipio del Distrito Metropolitano de Quito, 2020; 3 Prefeitura Municipal do Salvador, 2020; Municipalidad de Lima, 2021; Prefeitura da Cidade do Rio de 4 Janeiro, 2021; Prefeitura do Município de São Paulo, 2021). 5 6 Housing programs and initiatives that consider resilient construction, and site selection strategies, are still in 7 nascent stages (Martin et al., 2013). Initiatives in slum upgrading, social housing improvement and 8 regularizing land tenure, associated with infrastructure provision, do not usually focus on adaptation, 9 although they often focus on risk reduction. Those initiatives, associated with a housing policy that 10 guarantees access to land and decent housing, a comprehensive intervention in vulnerable neighbourhoods 11 for their adaptation to climate change, and CbA (community-based adaptation) strategies, including housing 12 self-management and the participation of cooperatives, shows the need and opportunity to move to an 13 transformative urban agenda that encompasses sustainable development, poverty reduction, disaster-risk 14 reduction, climate-change adaptation, and climate-change mitigation (high confidence) (Muntó, 2018; UN- 15 Habitat, 2018; Valadares and Cunha, 2018; Bárcena et al., 2020b; Núñez Collado and Wang, 2020; 16 Satterthwaite et al., 2020). 17 18 Several large cities are implementing municipal risk management plans and management and restoration 19 plans for hydrologically relevant areas, considering threats of drought and heat waves, integrated watershed 20 management and flood control programs (high confidence) (Municipio del Distrito Metropolitano de Quito, 21 2020; Prefeitura Municipal do Salvador, 2020; Municipalidad de Lima, 2021; Prefeitura da Cidade do Rio de 22 Janeiro, 2021; Prefeitura do Município de São Paulo, 2021). Quito and Rio de Janeiro are considered two 23 examples of comprehensive and effective city-level climate action that includes creating environment 24 protected areas, managing appropriate land use, household relocation and EWS in vulnerable to high- 25 precipitation areas associate to EbA, such as reforestation projects, to face natural hazards (ELLA, 2013; 26 Anguelovski et al., 2014; Calvello et al., 2015; Alcaldía de Quito, 2017; Sandholz et al., 2018; Prefeitura da 27 Cidade do Rio de Janeiro, 2021) (Section 12.6.1). EWS and the use of mapping tools experienced in La Paz 28 showed to be an effective adaptation measure facing increasing hydro-climatic extreme events (Aparicio- 29 Effen et al., 2018). 30 31 12.5.5.3.2 Green and grey infrastructure 32 Hybrid solutions, combining green and grey infrastructure (GGI), have been adopted for better efficiency in 33 flooding control (Ahmed et al., 2019; Drosou et al., 2019; Romero-Duque et al., 2020), sanitation, water 34 scarcity, landslide prevention and coastal protection (high confidence) (Section 12.5.6.4; Mangone, 2016; 35 Depietri and McPhearson, 2017; Leal Filho et al., 2018; McPhearson et al., 2018). The adoption of nature- 36 based solutions (NbS), which embraces well-known approaches such as green infrastructure (GI) and 37 ecosystem-based adaptation (EbA) (Pauleit et al., 2017; Le, 2020) has increased (Box 1.3). The Fund for the 38 Protection of Water (FONAG) and the Participative Urban Agriculture (AGRUPAR) are initiatives using 39 NbS in Quito (Section 12.6.1). Example of GGI is a stormwater detention pond, as a water storage solution 40 to flooding prevention, also allowing multiple uses of an urban space, adapting and revitalizing a degraded 41 area in Mesquita, Rio's metropolitan region (Jacob et al., 2019). These systemic and holistic solutions still 42 need to overcome governance and sectorial barriers to be more widely adopted (Herzog and Rozado, 2019; 43 Wamsler et al., 2020; Valente de Macedo et al., 2021). 44 45 Managing water in cities in an adaptive way has been central to reducing impacts such as floods and 46 contributes to water security (high confidence) (Van Leeuwen et al., 2016; Okumura et al., 2021). Many 47 cities facing frequent heavy storms that impact mostly underprivileged communities, slums and vulnerable 48 areas could benefit from the integrated NbS for disaster risk reduction and adaptation (high confidence) 49 (Sandholz et al., 2018; Ronchi and Arcidiacono, 2019). A study covering 70 Latin American cities estimates 50 that 96 million people would benefit from improving main watersheds with green infrastructure (Tellman et 51 al., 2018). In several municipal climate plans, NbS were introduced mainly to enhance rainwater 52 management, reduce energy consumption and urban heat areas, water quality, prevent landslides and offer 53 green areas (high confidence) (Gobierno de la Ciudad de Buenos Aires, 2020; Municipio del Distrito 54 Metropolitano de Quito, 2020; Prefeitura Municipal de Curitiba, 2020; Alcaldía de Medellín, 2021; 55 Municipalidad de Lima, 2021; Prefeitura da Cidade do Rio de Janeiro, 2021; Prefeitura do Município de São 56 Paulo, 2021). Sao Paulo's project for Jaguaré river proposes a large-scale landscape transformation applying 57 innovative multifunctional NbS instead of exclusively large, expensive and monofunctional hard engineered Do Not Cite, Quote or Distribute 12-71 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 solutions to manage stormwater (Marques et al., 2018; Herzog and Rozado, 2019). In Bogotá, the Humedales 2 foundation has restored wetlands to enhance areas near the reserve Van Der Hammen to improve water 3 quality and quantity, restore habitat for biodiversity, and provide flood protection (Portugal Del Pino et al., 4 2020). In Petrópolis, a medium-sized city in the hills of Rio de Janeiro state, the water service company has 5 implemented 10 NbS multifunctional micro wastewater treatment plants in low-income areas, helping to 6 reduce cascading impacts of storms, floods and epidemics (Herzog and Rozado, 2019). In Costanera Sur, 7 Buenos Aires, a public initiative to protect an auto-regenerated Plata riverbank, which had received 8 demolition material to create land, nowadays offers numerous ecosystem services for residents and attract 9 visitors activating the tourist industry and helping reducing riverine floods (Bertonatti, 2021; OICS, 2021). 10 11 Hybrid solution on water management that can merge traditional interventions on urban areas with 12 sustainable urban drainage systems (SUDS) (Davis and Naumann, 2017), considering small scale low-impact 13 development (LID) measures scattered over the watershed instead of concentrate huge hydraulic grey 14 structures, can help reduce the risk and damage of flooding (high confidence) (Miguez et al., 2014; Miguez 15 et al., 2015a; Depietri and McPhearson, 2017; Da Silva et al., 2018a; de Macedo et al., 2018). Quito's 16 climate plan explicitly cites the strategy for implementing blue and grey infrastructure to reduce risk due to 17 extreme precipitation and its associated impacts such as flooding and landslides and the possible impact of 18 water scarcity (Municipio del Distrito Metropolitano de Quito, 2020). The Integrated Iguaçu-Sarapuí River 19 Basin Flood Control Master Plan, in Rio's metropolitan area, combined different solutions for flood 20 protection, focusing on river restoration by retrofitting levee systems combined with adapting land use to 21 provide a multifunctional landscapes as an alternative to bring together green and grey solutions, composing 22 urban parks to prevent further paving and avoid irregular occupation of river banks and provide storage 23 capacity for damping flood peaks (Miguez et al., 2015b). 24 25 Many cities are implementing adaptation measures on integrated water and flood management systems 26 (Sarkodie and Strezov, 2019), improving basic sanitation services (medium confidence: medium evidence, 27 high agreement). Main strategies are established by NAPs recurrently focusing on improving water 28 distribution network and reservoir systems, as Honduras (Government of Honduras, 2018) and Ecuador 29 (Mills-Novoa et al., 2020), sewage and effluent treatment, as Guatemala, Brazil and Paraguay (Government 30 of Brazil, 2007; Government of Guatemala, 2016; Government of Paraguay, 2017), facing water scarcity and 31 environmental degradation. Local authorities follow this guideline as in the effort to maintain and upgrade 32 existing drainage systems in Georgetown (Mycoo, 2014), or in Medellin, focusing on improving drainage 33 systems to prevent landslides or flooding (Núñez Collado and Wang, 2020; Alcaldía de Medellín, 2021). Rio 34 de Janeiro has constructed three large stormwater detention reservoirs to deal with frequent flood, (Prefeitura 35 da Cidade do Rio de Janeiro, 2015), adopting a set of exclusively grey solutions, not combined to NbS that 36 could improve urban flood resilience (Rezende et al., 2019). The main proposed actions still consider the 37 traditional approach in improving the hydraulic capacity of urban drainage systems as an adaptive measure 38 (high confidence) (Gobierno de la Ciudad de Buenos Aires, 2020; Prefeitura Municipal do Salvador, 2020; 39 Municipalidad de Lima, 2021; Prefeitura da Cidade do Rio de Janeiro, 2021). In addition to this strategy, 40 several local plans propose actions for the retention and storage of rainwater, both in the urban drainage 41 network with a smaller intervention scale (Prefeitura Municipal de Curitiba, 2020), as well as along rivers 42 and canals with large-scale works (medium confidence: medium evidence, high agreement) (Gobierno de la 43 Ciudad de Buenos Aires, 2020; Prefeitura Municipal de Curitiba, 2020; Alcaldía de Medellín, 2021; 44 Prefeitura da Cidade do Rio de Janeiro, 2021). 45 46 12.5.5.3.3 Mobility and transport system 47 Mobility and transport systems have a key role in urban resilience (high confidence) (Walker et al., 2014a; 48 Caprì et al., 2016; Espinet et al., 2016; Lee and Lee, 2016; Ford et al., 2018; Mehrotra et al., 2018; Quinn et 49 al., 2018). Examples reported in scientific literature assessed are focusing on mitigation strategies even when 50 labelled as adaptation (da Silva and Buendía, 2016; Di Giulio et al., 2018; Valderrama et al., 2019; Goes et 51 al., 2020). 52 53 The integration of transport and land use planning and the improvement of public transport, also as important 54 mitigation actions, appears as a consensus in countries' adaptation plans, nevertheless the emphasis on 55 mobility and transport systems on the many NAP published is low (medium confidence: medium evidence, 56 high agreement). Honduras, Costa Rica and El Salvador's NAP are not approaching adaptation or mitigation 57 in the sector, while Peru, Ecuador, Guatemala and Paraguay ones focus on mitigation only. Chile, Colombia Do Not Cite, Quote or Distribute 12-72 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 and Brazil's NAP focus on both mitigation and adaptation of mobility and transport systems. Chile and 2 Colombia's plans dedicated specific action lines to adapt mobility and transport systems to climate change, 3 whilst Brazil published a NAP's complementary volume dedicated exclusively to the sectoral strategies, 4 although presents only general guidelines (Government of Peru, 2010; Government of Chile, 2014; 5 Government of Ecuador, 2015; Government of Brazil, 2016; Government of Colombia, 2016; Government 6 of Guatemala, 2016; Government of Paraguay, 2017; Government of Costa Rica, 2018; Government of 7 Honduras, 2018; Government of El Salvador, 2019). 8 9 In municipal scale, assessing the biggest cities, São Paulo, Rio de Janeiro Lima and Santiago stands out for 10 including mobility and transport as one of the strategic axes of its climatic plans, but yet prioritizing 11 mitigation, while Buenos Aires and Bogotá do not deepen the issue in their plans (Gobierno de la Ciudad de 12 Buenos Aires, 2015; Prefeitura da Cidade do Rio de Janeiro, 2016; Alcaldía Mayor de Bogotá D.C., 2018; 13 Municipalidad de Lima, 2021; Municipalidad de Santiago, 2021; Prefeitura do Município de São Paulo, 14 2021). Most of those same cities have sectoral mobility plans, which are key tools to urban resilience. Those 15 plans, however, do not focus on adaptation actions, although emphasizing mitigation (Government of Peru, 16 2005; Gobierno de la Ciudad de Buenos Aires, 2011; Prefeitura do Município de São Paulo, 2015; Alcaldía 17 Mayor de Bogotá D.C., 2017; Ilustre Municipalidad de Santiago, 2019; Município de Rio de Janeiro, 2019). 18 19 12.5.6 Health and Wellbeing 20 21 The most common adaptation strategies include the development of climate services such as epidemic 22 forecast tools, integrated climate-health surveillance and observatories and forecasting climate-related 23 disasters (floods, heat waves). GIS technologies are being used to identify locations where vulnerable 24 populations are exposed to climate hazards and associated health risks. 25 26 12.5.6.1 Climate services for health 27 28 The measures most directly linked to diminishing risk are those related to climate services for health (high 29 confidence). Climate services provide tailored, sector-specific information from climate forecasts to support 30 decision making (WHO and WMO, 2016); they allow decision makers and practitioners to plan 31 interventions in anticipation of a weather/climate event (Mahon et al., 2019). More recently, climate 32 services, such as early warning systems (EWS) and forecast models, have been promoted for the health 33 sector (WHO and WMO, 2012; WMO, 2014; WHO and WMO, 2016; Thomson and Mason, 2018) and are 34 an important adaptation measure to reduce the impacts of climate on health (high confidence). To guide this 35 process, the Global Framework for Climate Services (GFCS) issued a Health Exemplar (Lowe et al., 2014; 36 WMO, 2014) which aims for stakeholder engagement between health and climate actors at all levels to 37 promote the effective use of climate information within health research, policy and practice. 38 39 There exist at least 24 EWS in SA to avoid deaths and injuries from floods in the countries such as 40 Argentina, Colombia, Ecuador, Bolivia, Brazil, Peru, Uruguay and Venezuela (Bravo et al., 2010; Bidegain, 41 2014; Moreno et al., 2014; Dávila, 2016; del Granado et al., 2016; López-García et al., 2017; Carrizo Sineiro 42 et al., 2018). A total of 149 emergency prevention and response systems are reported in CA (UNESCO, 43 2012). In addition, some countries implement programs for the relocation of families who are in risk 44 condition, like in Bogotá and Medellin, Colombia (World Bank, 2014; Watanabe, 2015). 45 46 Epidemic forecast tools are an example of an adaptation measure being developed and/or implemented in 47 this region (high confidence). Climate-driven forecast models have been developed for dengue in Ecuador, 48 Puerto Rico, Peru, Brazil, Mexico, Dominican Republic, and Colombia (Lowe et al., 2013; Eastin et al., 49 2014; Johansson et al., 2016; Lowe et al., 2017; Johansson et al., 2019); for Zika virus infections across the 50 Americas (Muñoz et al., 2017); for cutaneous leishmaniasis in Costa Rica and Brazil (Chaves and Pascual, 51 2006; Lewnard et al., 2014); for Aedes-borne diseases across the Americas (Muñoz et al., 2020b); and a 52 nowcast model for chikungunya virus infections across the Americas (Johansson et al., 2014). In Ecuador, a 53 prototype system utilized forecasts of seasonal climate and ENSO forecasts of to predict dengue 54 transmission, providing the health sector with warnings of increased transmission several months ahead of 55 time (Stewart-Ibarra and Lowe, 2013; Lowe et al., 2017). Despite these advances, few tools have become 56 operational and mainstreamed in decision making processes. However, Brazil and Panama have been able to Do Not Cite, Quote or Distribute 12-73 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 operationalize an early warning system for the surveillance of dengue fever transmission (Codeço et al., 2 2016; McDonald et al., 2016). 3 4 One of the most promising climate services for the health sector are heat and cold early warning and alert 5 systems (medium confidence). These have been developed by the national meteorological institutes in Peru, 6 Argentina, and Uruguay (Bidegain, 2014). A heat alert system was implemented in Argentina in 2017 and 7 daily alerts are issued for 57 localities across the country. A stoplight colour scheme is used to issue alerts, 8 identifying specific groups at risk and actions to be taken to reduce the risk (Herrera et al., 2018b). 9 10 The public dissemination of climate-health warnings via bulletins, websites, and other outlets can be an 11 adaptation measure to climate change and weather variability to diminish health risk (high confidence). The 12 information produced is systematized to be communicated to the authorities and general public. The 13 Caribbean Health-Climatic Bulletin has been issued quarterly since 2018 to health ministries across the 14 region, including CA and NSA. Regional climate and health authorities meet to review 3 month climate 15 forecasts and issue statements about the probable impacts on health (Trotman et al., 2018). In Panamá, 16 information on dengue is distributed in a monthly bulletin that is used by health authorities to inform vector 17 control activities (McDonald et al., 2016). Another example was the climate-driven forecast of dengue risk 18 that was produced prior to Brazil's 2014 FIFA World Cup to inform disease prevention interventions (Lowe 19 et al., 2014; Lowe et al., 2016). In Colombia, the Intersectoral National Technical Commission for 20 Environmental Health publishes a monthly bulletin with regional weather forecast and potential effects on 21 health (CONASA, 2019). Paraguay improves epidemiological surveillance and trains first level health staff 22 via information campaigns on the prevention of climate sensitive diseases, and promotes health networks 23 with the participation of civil society (Environmental Secretariat of Paraguay, 2011). 24 25 12.5.6.2 Integrated climate-health surveillance and observatories 26 27 Integrated health-climate surveillance systems are another key adaptation strategy (medium confidence). This 28 information can be used by the health sector to inform decision making about when and where to deploy a 29 public health intervention. It can also feed into an EWS, particularly if the data are compatible in format and 30 spatiotemporal scales. An integrated health-climate surveillance system for vector borne disease control was 31 developed in southern coastal Ecuador through a partnership among the climate and health sectors and 32 academia (Borbor-Cordova et al., 2016; Lowe et al., 2017). Additionally, an interdisciplinary multinational 33 team working at the border of Ecuador and Peru created a cooperation network for climate-informed dengue 34 surveillance (Quichi et al., 2016) and successful binational collaboration resulted in the local elimination of 35 malaria (Krisher et al., 2016). Similar is the innovative community-based data collection to understand and 36 find solutions to rainfall-related diarrheal diseases in Ecuador (Palacios et al., 2016). 37 38 Climate and health observatories are a promising strategy being developed at subnational, national (e.g., 39 Brazil, Argentina) and regional levels (high confidence) (Muñoz et al., 2016; Rusticucci et al., 2020). The 40 Brazilian Observatory of Climate and Health brings together climate and health information for the Amazon 41 region of Manaus (Barcellos et al., 2016). At a national level, Brazil has created the climate and health 42 observatory, where information and data visualizations are available for various climate-sensitive health 43 indicators (Ministério da Saúde and FIOCRUZ, 2021). 44 45 12.5.6.3 Vulnerability and risk maps 46 47 Vulnerability and risk maps have been widely used as an adaptation strategy to understand the potential 48 impacts of climate on health outcomes both directly (e.g., maps of disease risk) and indirectly (e.g., maps of 49 populations vulnerable to climate disasters) (high confidence). There are many examples where climate 50 services have been used to construct vulnerability maps for health outcomes, including maps in the 51 aforementioned Climate-Health Observatories. Dengue, malaria, and Zika vulnerability maps using climate, 52 social, and environmental information has been developed in Brazil and Colombia (Cunha et al., 2016b; 53 López-Álvarez, 2016; Pereda, 2016; IDEAM, 2017). Argentina is focused on improving the health system by 54 using Climate Change Risk Map System as a tool that identifies the risks and allows assessing their 55 management (OPS and WHO, 2018). 56 Do Not Cite, Quote or Distribute 12-74 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Vulnerability and risk maps for climate disasters have been developed at the city level, for example in 2 Bogotá, Cartagena de Indias, and Mocoa in Colombia (Yamin et al., 2013; Guzman Torres and Barrera 3 Arciniegas, 2014; Tehelen and Pacha, 2017; Zamora, 2018); and for the metropolitan district of Quito in 4 Ecuador (Tehelen and Pacha, 2017). In addition, vulnerability maps were created for the primary road 5 network of Colombia (Tehelen and Pacha, 2017). At the regional level, vulnerability maps using climate 6 change probability, disaster risk and food insecurity variables has been produced for the Andean region 7 (WFP, 2014). In Brazil, vulnerability maps considering exposure, sensitivity, and adaptive capacity, coupled 8 to climate scenarios, were designed to support the National Adaptation Plan on a municipal scale (Chang and 9 Garcia, 2018; Duval et al., 2018; Marinho and Silva, 2018; Menezes, 2018; Santos and Marinho, 2018; Silva 10 et al., 2018). A Climate Change Vulnerability Index was used to generate vulnerability maps for countries of 11 Latin American and Caribbean region (Vörösmarty et al., 2013; CAF, 2014). 12 13 12.5.6.4 Other adaptation actions 14 15 Diverse adaptation measures are being implemented through public policies, private households' responses, 16 or communal management that directly or indirectly reduce the impacts of climate change on human health 17 (high confidence) (Table 12.9). Private and communal management measures could be considered indirect 18 measures, because they might be adopted even in the absence of climate change. 19 20 21 Table 12.9: Hazards from climate change that impact human health and examples of adaptation strategies proposed or 22 implemented in CSA. Based in McMichael et al. (2006); Miller et al. (2013a); Miller et al. (2013b); Miller et al. 23 (2013c); Miller et al. (2013d); Hardoy et al. (2014); IPCC (2014); Janches et al. (2014); Lee et al. (2014); Mejia (2014); 24 Sosa-Rodriguez (2014); Vergara et al. (2014); Lemos et al. (2016); Villamizar et al. (2017); Magoni and Munoz (2018); 25 Zhao et al. (2019). Hazard and Examples of adaptation strategies impacts on human health Public Private Communal Extreme heat and cold: · Creation of urban green · Cooling by swamp · Training of deaths / illness by spaces/ coolers, air community health thermal stress conditioning, open volunteers to · Health promotion windows, wet the recognize and treat campaigns. floors, shade trees. heat strain. · Establish shelters during · Bioclimatic building heat waves design · Technology transfer for home heating Extreme rainfall, · Early warning systems · Green-grey · Communal efforts to wildfire, wind speed: (EWS) for extreme climate infrastructure to clear debris from injury / deaths from events. prevent landslides. canals to reduce floods, storms, flood risk cyclones, bushfires and · Safe housing programs and · Insurance mechanisms landslides (Key risk 2, relocation and financing for long- · Cooperative efforts Table 12.6). term recovery. to rebuild following a · Green-grey infrastructure flood event (e.g., channels, drainage systems) Drought and dryness: · Formalizing land · Water infrastructure · Incorporation of local poor nutrition due to ownership for small and irrigation. stakeholders in reduced food yields farmers and Indigenous formulating and dehydration due to people. · Soil moisture retention adaptation responses. limited or inadequate techniques management of · Address emerging water · Recognition of freshwater (Key risk 1, conflicts. · Insurance mechanisms. Indigenous and local Table 12.6). · Selection of drought wisdom and knowledge. resistant crops. Do Not Cite, Quote or Distribute 12-75 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report Changes in climate · Restoration of watersheds · Water disinfection: · Participatory water that promote microbial · Integrated health-climate boiling, chlorination. management proliferation: food strategies, including poisoning, and unsafe surveillance · Purchasing water or protection of drinking water (Key · Improve access to drinking water filters. drinking water risk 3, Table 12.6). sources. water, drainage, sanitation and waste removal. Changes in climate · Vector control · Use of bed nets and · Community that affect vector- · EWS for epidemics screens volunteers to collect pathogen host relations · Nature-based solutions blood smears for and infectious disease · Use of repellent and malaria diagnosis geography/seasonality (NbS) (e.g., forest insecticides. (Key risk 4, Table conservation) · Community-led 12.6). · Elimination of standing elimination of vector water. habitat. Sea level rise and · Improve governance of · Improve water · Incorporation of local storm surges: impaired water utilities. efficiency in stakeholders in crop, livestock and agriculture. formulating fisheries yields; unsafe · Address emerging water adaptation responses. drinking water, leading conflicts. to impaired nutrition · Recognition of (Key risk 8, Table · Protection, restoration and Indigenous and local 12.6). soil conservation to wisdom and recharge aquifers. knowledge. Environmental · Long-term risk · Identification of · Community-led degradation: loss of management planning for alternative livelihoods. efforts to reforest and livelihoods and cities. restore/protect displacement leading watersheds. to poverty and adverse · Sustainable forestry health outcomes programs. (related to Key risk 6, Table 12.6). · Protection and restoration of lacustrine areas. 1 2 3 Participatory management can be relevant in the case of mosquito-borne disease prevention (e.g., dengue 4 fever or malaria), where the reduction in mosquito habitat in one area or `hot spot' can reduce the risk for all 5 surrounding households. This approach is also relevant when considering new places where vector-borne 6 diseases can emerge because of changes in climate (Andersson et al., 2015). 7 8 Adaptation strategies implemented by the public sector include a diverse suite of strategies ranging from 9 creation of green spaces in urban areas, relocation of families located in disaster prone areas, ecosystem 10 restoration, improved access to clean water, among many others (high confidence) (Table 12.9). Building 11 green-grey infrastructure (GGI) has been a popular public adaptation measure to reduce deaths and injuries 12 because of floods (Section 12.5.5.3.2). Infrastructure has been improved at schools, public buildings and 13 drainage systems in cities such as Bogota, Colombia (World Bank, 2014) and La Paz, Bolivia (Fernández 14 and Buss, 2016). In Brazil, channel works were implemented to reduce the flooding of the Tiete River, 15 which crosses the metropolitan area of Sao Paulo; these projects were designed based on simulated flood 16 scenarios (Hori et al., 2017). 17 18 Another example of a public adaptation measure is protection and restoration of natural areas, which have 19 the potential to decrease the transmission of water- and vector-borne infectious diseases (medium 20 confidence: robust evidence, low agreement). Studies have shown that these measures can diminish the cases 21 of malaria and diarrhoea in Brazil, and cases of diarrhoea in children in Colombia (Bauch et al., 2015; 22 Herrera et al., 2017; Chaves et al., 2018). However, deforestation and malaria have a complex relationship 23 that relies on local context interactions, where land use and land cover change present an important role due 24 to vector ecology alterations and social conditions of human settlements (Rubio-Palis et al., 2013). Forest 25 conservation can improve hydrological cycle control and soil erosion that can help to improve water quality 26 and reduce the burden of water-borne diseases. In addition, forest cover can help to diminish the habitat for 27 larval mosquitoes that transmit malaria. These measures can help to design policies in sites where these Do Not Cite, Quote or Distribute 12-76 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 problems do not currently exist but can emerge as a consequence of climate change and the increase in the 2 frequency of weather extreme events. 3 4 12.5.6.5 Challenges and opportunities 5 6 Despite the proliferation of disaster EWS in the region, only 37 can be considered operational, because many 7 of these systems are not operating or functioning properly, or do not meet the requirements to be considered 8 EWS (UNESCO, 2012). Sustainable financing and political support are needed to ensure the functioning of 9 disaster EWS (high confidence) (Table 12.11). Several studies identified difficulties in implementing disaster 10 EWS due to a lack of community engagement and response to the alerts that are issued (del Granado et al., 11 2016; López-García et al., 2017). To address these challenges, the document "Developing Early Warning 12 Systems: A Checklist" provides guidance for the implementation of a people centred approach to early 13 warning systems as proposed in the Hyogo Framework for Action 2005­2015 (Wiltshire, 2006). 14 15 With respect to the development of climate-driven epidemic forecasts, efforts are needed to improve the 16 utility of such forecasts for the health sector. Few such forecasts have been operationalized to inform health 17 sector decision making. A review of 73 studies that predicted and forecasted Zika virus infections (42% from 18 the Americas) identified a high degree of variation in access, reproducibility, timeliness, and incorporation of 19 uncertainty (Kobres et al., 2019). A recent systematic review of epidemic forecasting and prediction studies 20 found that no reporting guidelines exist; the development of guidance to improve transparency, quality and 21 implementation of forecast models in the public health sector was recommended (Pollett et al., 2020). An 22 earlier review of dengue early warning models found that few models incorporated both spatial and temporal 23 aspects of disease risk (Racloz et al., 2012), limiting their potential application as an adaptation strategy by 24 the health sector. Advances have been made in the last decade with respect to modelling and computing 25 tools, increasing access to digital climate information and health records, and the use of earth observations to 26 forecast climate sensitive diseases (Fletcher et al., 2021; Wimberly et al., 2021). 27 28 The growing field of implementation science ­defined as "a discipline focused on systematically examining 29 the gap between knowledge and action"­ is another opportunity to address the challenges and barriers to 30 using climate information for health sector decision making (Boyer et al., 2020). Implementation science in 31 the health sector in CSA is nascent; research in this area could help to address barriers to mainstreaming 32 climate information in the health sector as an adaptation strategy (Table 12.11; Table SM12.7). 33 34 12.5.6.6 Governance and Financing. 35 36 A description of the governance and financing dimensions of the feasibility of implementing EWS is 37 presented in Table 12.11 and Table SM12.7. 38 39 12.5.6.6.1 National Health Plans 40 Some countries have developed national plans on health including the role of climate. Chile has a Climate 41 Change Adaptation Plan of the Health Sector that proposes several actions to enhance monitoring, 42 institutions and citizens information and education (Ministry of Health of Chile and Ministry of Environment 43 of Chile, 2016). Based on the identification of vulnerability to climate change, Colombia has developed 44 eleven regional adaptation plans to strengthen institutional capacities; climate change education for 45 behavioural changes; and cost estimation to promote health resilience (WHO and UNFCCC, 2015). In 46 addition, El Salvador implemented actions to strengthen health infrastructure through high latrines for 47 housing in flood communities, as well as other measures focused on water supply and quality based on an 48 education and awareness program (Ministry of Environment and Natural Resources of El Salvador, 2013). 49 Only Brazil and Peru have implemented actions so far in the region derived from national health adaptation 50 plans, and only Brazil completed a national assessment of impacts, vulnerability and adaptation for health 51 (Watts et al., 2018). Some countries include health as a priority sector in their National Adaptation Plans, as 52 is the case of Ecuador, and Costa Rica, which has a national plan addressing the prevention and care of 53 climate-sensitive diseases coupled to a National Health Plan (2016-2020) (Ministry of Health Costa Rica, 54 2016; Jiménez, n.d.). 55 Do Not Cite, Quote or Distribute 12-77 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 12.5.6.6.2 National Disaster Management Plans 2 National Risk Management Plans or National Disaster Response Plans are tools for adapting to climate 3 change that can help to diminish death and injuries because of disasters (high confidence). These Plans are 4 generally promoted by governments as national instruments that guide the processes of estimating, 5 preventing and reducing disaster risk. An updated national risk management plans has been found for 6 Guatemala (CONRED, 2014), Honduras (COPECO, 2014), El Salvador (Ministry of Health of El Salvador, 7 2017), Costa Rica (CNE, 2016), Ecuador (SGR, 2018), Peru (SGRD et al., 2014), Argentina (Ministerio de 8 Seguridad de Argentina, 2018), Bolivia (VIDECI, 2017), Chile (ONEMI, 2015) and Colombia (UNGRD, 9 2015). It has been shown in Brazil that information on drought conditions can be used to reduced health 10 impacts of drought using a national disaster risk reduction framework (Sena et al., 2016). 11 12 12.5.7 Poverty, Livelihood and Sustainable Development 13 14 Climate change impacts are increasing and exacerbating poverty and social inequalities, affecting those 15 already vulnerable and disfavoured, generating new and concatenated risk challenging climate resilient 16 development pathways (high confidence) (Section 8.2.1.4; Shi et al., 2016; Otto et al., 2017; Johnson et al., 17 2021) (). Poverty, high levels of inequalities and pre-existing vulnerabilities also can be worsened by climate 18 change policies (Antwi-Agyei et al., 2018; IPCC, 2018; Roy et al., 2018; Eriksen et al., 2021). Those already 19 suffering are losing their livelihoods and reducing their development options; poor populations and countries 20 are more vulnerable and have lower adaptive capacity to climate change (very high confidence) (Section 21 8.5.2.1; Rao et al., 2017). 22 23 Inequality is growing, being a CSA structural characteristic; Gini index average for Latin American 24 countries (including Mexico) was decreasing to 0.466 in 2017, where 1% richest got 22 times more income 25 than 10% poorest (ECLAC, 2019b; Busso and Messina, 2020), but in 2018, 29.6% of Latin America were 26 poor population (increased to 182 million) and extreme poverty 10.2%; in 2018 (increased to 63 million) 27 (ECLAC, 2019b) and in 2020, due to COVID crisis, Gini coefficient projection of increases are ranging from 28 1.1% to 7.8% (ECLAC and PAHO, 2020), poverty increased to 33.7% (209 millions) and extreme poverty to 29 12.5% (78 millions) (ECLAC and PAHO, 2020; ECLAC, 2021). Those poverty and extreme poverty rates 30 are higher among children, young people, women, Indigenous Peoples (Reckien et al., 2017; Busso and 31 Messina, 2020), migrant (Dodman et al., 2019) and rural population. Climate change impacts in 32 differentiated ways, even within a household there may be important differences in relation to age, gender, 33 health and disability; these factors may intersect with one another (high confidence) (Reckien et al., 2017; 34 Busso and Messina, 2020). 35 36 In IPCC's TAR, AR4 and AR5, WG II recognized higher risks associated with poor living conditions, 37 substandard housing, inadequate services, location in hazardous sites due to no alternatives and the need to 38 work more strongly on strengthening governance structures involving residents, community organizations 39 amongst others (Wilbanks et al., 2007; Revi et al., 2014). The AR5 CSA chapter stated that poverty levels 40 remained high (45% for CA and 30% for SA in 2010) despite years of sustained economic growth. Poor and 41 vulnerable groups are disproportionately affected in negative ways by climate change (Section 8.2.1.4; 42 Section 8.2.2.3; SR15 Section 5.2 and Section 5.2.1, Roy et al., 2018) ) due to physical exposure derived 43 from the place where they live or work, illiteracy, low income and skills, political and institutional 44 marginalization tied to lack of recognition of informal settlements and employments, poor access to good 45 quality services and infrastructure, resources, information, and other factors (very high confidence) (UN- 46 Habitat, 2018; SR15 Sections 5.2.1, 5.6.2, 5.6.3, 5.6.4, Roy et al., 2018). 47 48 International agreements aim for climate resilient development pathways where efforts to eradicate poverty, 49 reduce inequalities and promote fair and cross-scalar adaptation and mitigation are strengthened. The 50 Sustainable Development Goals (SDG) first and second objectives aim to reduce poverty leaving no one 51 behind (UN General Assembly, 2015). Although researchers argue that poverty is mischaracterized having 52 multiple dimensions (Castán Broto and Bulkeley, 2013) (Section 8.1.1), that biodiversity loss, climate 53 change and pollution will undermine efforts on 80% of assessed SDG targets, that biodiversity and climate 54 change must be tackle together (Pörtner et al., 2021; United Nations Environment Programme, 2021) and 55 LAC countries due to COVID crisis have uneven SDG progress (high confidence) (ECLAC, 2020). 56 Do Not Cite, Quote or Distribute 12-78 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 12.5.7.1 Challenges and Opportunities 2 3 Climate change exacerbates pre-existing conditions and moving in the opposite direction in the search for 4 resilience, equity and sustainable development (Tanner et al., 2015b; Bartlett and Satterthwaite, 2016; 5 Kalikoski et al., 2018; Bárcena et al., 2020a). Existing inequalities in the provision and consumption of 6 services are bound to be exacerbated by future risks and uncertainties associated with climate change 7 scenarios (Miranda Sara et al., 2017). Climate change will be a major obstacle in reducing poverty (high 8 confidence) (Bartlett and Satterthwaite, 2016; Allen et al., 2017a; Hallegatte et al., 2018; UN-Habitat, 2018; 9 United Nations Environment Programme, 2021), even affecting wealthier populations that become 10 vulnerable facing climate change scenarios (WGI AR6 Chapter 12, Ranasinghe et al., 2021), dragging them 11 into poverty, erasing decades of work and asset accumulation. 12 13 CSA is highly urbanized, the poor vast majority live in urban areas (except in Central America) while urban 14 extreme poverty is becoming more relevant (Rosenzweig et al., 2018; Dodman et al., 2019; Almansi et al., 15 2020; Sette Whitaker Ferreira et al., 2020), with those living in informal settlements and working within 16 informal economy are critical on each city's economy (Satterthwaite et al., 2018; Satterthwaite et al., 2020). 17 Many households in the region's cities live in precarious neighbourhoods with insufficient infrastructure and 18 substandard housing (Adler et al., 2018; Rojas, 2019). On average, between 21% and 25% of the urban 19 population lives in informal settlements (Jaitman, 2015; UN-Habitat, 2015; Rojas, 2019; Sandoval and 20 Sarmiento, 2019). This hides important disparities: Habitat III reports by individual countries the percentage 21 of urban population living in informal settlements ranged from 5% to 60%; in absolute terms 105 million 22 people living in precarious conditions (106 million estimated in 1990) (Section 12.5.5; Sandoval and 23 Sarmiento, 2019). 24 25 High levels of inequality and informality remain the biggest challenges for adaptation measures being 26 effective (Rosenzweig et al., 2018; Dodman et al., 2019). The interaction of projected impacts with existing 27 vulnerabilities in the region (as hunger, malnutrition and health inequalities, arising from its social, economic 28 and demographic profile), affect CSA development and well-being in different ways (Reyer et al., 2017) 29 increasing poverty and inequality risking the paths for sustainable development (Section 18.1.1; Reckien et 30 al., 2017). 31 32 The uneven enforcement of land-use regulations, relocations and evictions on behalf of environmental risk 33 management and climate adaptation is contested (Brockington and Wilkie, 2015; Lavell, 2016; Quimbayo 34 Ruiz and Vásquez Rodríguez, 2016a; Quimbayo Ruiz and Vásquez Rodríguez, 2016b; Anguelovski et al., 35 2018; Anguelovski et al., 2019; Shokry et al., 2020; Chávez Eslava, 2021; Oliver-Smith, 2021). This points 36 to caution in framing climate adaptation and resilience related interventions as equally benefiting everyone 37 (high confidence) (Brown, 2014; Chu et al., 2016; Connolly, 2019; Romero-Lankao and Gnatz, 2019; 38 Johnson et al., 2021) and the need for incorporating equality and justice dimensions (very high confidence) 39 (Section 18.1.2.2; Agyeman et al., 2016; Meerow and Newell, 2016; Romero-Lankao et al., 2016; Shi et al., 40 2016; Reckien et al., 2017; Leal Filho et al., 2021) (). 41 42 Poor rural households in marginal territories with low productive potential and/or far away from markets and 43 infrastructure are highly vulnerable to climate change and easily fall into poverty-environment traps (high 44 confidence) (Barbier and Hochard, 2019; Heikkinen, 2021). Climate change is one of the main threats to 45 rural livelihoods in Central America, being agriculture a pillar for rural economies and food security, 46 especially for the poorest sectors that rely on subsistence crops in areas with low soil fertility and rainfall 47 seasonality (Bouroncle et al., 2017). 48 49 Impacts are likely to occur simultaneously, exacerbating those of the poorer but also creating new groups at- 50 risk (Miranda Sara et al., 2016; Rosenzweig et al., 2018; Dodman et al., 2019). The material basis for poor 51 and vulnerable urban and rural populations' adaptation are in a critical state across the CSA region, 52 magnifying extreme events' impacts, making CSA less resilient. The consequences in terms of social 53 vulnerability and livelihood will be widely felt, insofar the security and protection of critical assets (housing, 54 infrastructure, services - water, land and ecosystem services) continues to lay behind. Small businesses are 55 usually conducted within the same home and if the house is affected so is the business (Stein and Moser, 56 2015) adding another layer of vulnerability for them. 57 Do Not Cite, Quote or Distribute 12-79 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 As productivity declines, they seek outside income generation opportunities and rely on resource extraction 2 for subsistence and as an income generation activity, further increasing their vulnerability to climate change 3 (Barbier and Hochard, 2018a). Cycles of declining productivity, environmental degradation, wildlife 4 poaching and trafficking, search of outside employment, reduced incomes, livelihood opportunities and 5 poverty have been registered in rural El Salvador, Honduras, Amazonia (López-Feldman, 2014; Graham, 6 2017; Barbier and Hochard, 2018a). The protection of communities that defend and are dependent on 7 wildlife and natural environments requires immediate attention. In Latin America there are 8 million forest- 8 dependent people which represents about 82% of the region's rural extreme poor (FAO and UNEP, 2020). 9 10 Poverty and disaster risk reduction interlinked with climate change adaptation share a focus on identifying 11 and acting on local risks and their root causes, even having different lenses through which to view risk (very 12 high confidence) (IPCC, 2014; Allen et al., 2017a; Satterthwaite et al., 2018; UN-Habitat, 2018; 13 Satterthwaite et al., 2020). Construction of climate knowledge and risk perceptions affect decision-making to 14 define implementation priorities; the poor are less able to cope and to adapt avoiding "adaptation injustices" 15 (high confidence) (Mansur et al., 2016; Miranda Sara et al., 2017; Reckien et al., 2017; Hardoy et al., 2019). 16 17 Adaptation, social policies, poverty reduction and inequality are weakly articulated to daily or chronic risk 18 reduction. Poor residents are often caught in `risk traps', accumulated cycles of everyday risks and small- 19 scale disasters (medium confidence: medium evidence, high agreement) (Bartlett and Satterthwaite, 2016 ; 20 Mansur et al., 2016; Allen et al., 2017a; Leal Filho et al., 2021), being exacerbated by climate risks and 21 COVID pandemic with the most vulnerable populations suffering. Chronic and every day risks (poor access 22 to infrastructure, services, incomes, housing, tenure, education, security, location and poor-quality 23 environment, networks and having a voice) are often exacerbated and generate new unknown risks by 24 climate change (medium confidence: medium evidence, high agreement) (Bartlett and Satterthwaite, 2016; 25 Mansur et al., 2016; Satterthwaite et al., 2018; Leal Filho et al., 2021), extreme events and risks related to 26 ENSO oscillation. All these risks need to be considered simultaneously (UN-Habitat, 2018). Risks are 27 seldom distributed equally highlighting socioeconomic inequalities and governance failures (high 28 confidence) (IPCC, 2014; Bartlett and Satterthwaite, 2016; Rasch, 2016; Romero-Lankao et al., 2018). 29 30 Adaptation, disaster risk reduction together with social and poverty reduction policies contribute to 31 sustainable development (Hallegatte et al., 2018; Satterthwaite et al., 2020), and improve prospects of 32 climate resilient pathways (Section 18.1.1). Without pro-poor interventions, adaptation options could 33 reinforce poverty cycles (Kalikoski et al., 2018). Secure locations, good quality infrastructure, services and 34 housing are critical to reduce risks from extreme climate events (Satterthwaite et al., 2018; Dodman et al., 35 2019). 36 37 12.5.7.2 Governance and Finance 38 39 Poor and most vulnerable groups evidence limited political influence, fewer capacities and opportunities to 40 participate in decision and policy making, are less able to leverage government support to invest on 41 adaptation measures linked with poverty, inequality and vulnerability reduction (very high confidence) 42 (Chapter 8; Miranda Sara et al., 2017; Reyer et al., 2017; Kalikoski et al., 2018; Dodman et al., 2019; 43 Satterthwaite et al., 2020). 44 45 Existing unbalances on power relations, corruption, structural historic problems and high levels of risk 46 tolerance (Miranda Sara et al., 2016) constitute climate governance barriers for implementing more effective 47 adaptation and preventive measures. Corruption, particularly in the construction and infrastructure sector, 48 has proven to be a barrier for CSA development even reproducing and reconstructing risks (French and 49 Mechler, 2017; Vergara, 2018; Durand, 2019). Critical infrastructure and valuable assets continue to be 50 placed in vulnerable areas (Calil et al., 2017; Escalante Estrada and Miranda, 2020) evidencing the 51 persistence of maladaptation and adaptation deficit (Villamizar et al., 2017). 52 53 Social organization, participation and governance reconfiguration are essential for building climate resilience 54 (very high confidence) (Stein and Moser, 2015; Kalikoski et al., 2018; Satterthwaite et al., 2018; Stein et al., 55 2018; Hardoy et al., 2019; Stein, 2019; Satterthwaite et al., 2020; Miranda Sara, 2021). Adaptation measures 56 have trade-offs that need to be acknowledged and acted upon, most importantly by developing the capacity 57 to convene discussions that draw in all key actors and commit them to do things differently (Almeida et al., Do Not Cite, Quote or Distribute 12-80 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2018; Hardoy et al., 2019). Collaborative approaches integrating groups and organizations (e.g., saving, 2 women's groups, clubs, vendor associations, cooperatives) contributing to the exchange of information, to 3 visibilize people's needs, to generate safety networks, and to negotiate for improvements and enhance 4 adaptive capacity. 5 6 12.5.7.3 Adaptation options 7 8 Effective adaptation can be achieved by addressing pre-existing development deficits, particularly the needs 9 and priorities of informal settlements and economies (Revi et al., 2014; UN-Habitat, 2018). There is urgency 10 for social systems to better respond to climate related risks and increase their adaptive capacity (Lemos et al., 11 2016) focus on path dependency, lock ins and poor specific needs (Leal Filho et al., 2021). 12 13 The linkages between climate adaptation and poverty are not clearly addressed at national level (Kalikoski et 14 al., 2018). A revision of some NDCs presented by CSA countries (https://unfccc.int), shows that NDCs are 15 developed with almost no connection to poverty and livelihoods. Exceptions include Bolivia whose NDC 16 developed the "Good life" concept, as an alternative development pathway, supporting sustainable 17 livelihoods as a mean to eradicate poverty; Honduras asserts that climate action should improve living 18 conditions; Peru defined a poverty and vulnerability reduction approach and El Salvador conditioned its 19 NDCs to macroeconomic stability, economic growth and poverty reduction. A sustainable development 20 approach permeates in proposed actions for sectors as energy, agriculture, transport, water, and forestry. 21 22 Adaptive capacity is linked to addressing climate related risks (specific capacity) and structural deficits 23 (generic capacity), synergies and a strategic balance between both is necessary (Eakin et al., 2014; Lemos et 24 al., 2016). Adaptation institutional context can undermine one form of capacity with repercussions on the 25 other compromising overall adaptation and sustainable development (Eakin et al., 2014). 26 27 Literature assessing the effectiveness of pro-poor or community based adaptation practices and livelihood 28 options continues to be weak, even though are increasingly documented, as in AR5 (Magrin et al., 2014). 29 Great variety of measures are being applied, financial instruments to strengthen and protect livelihoods and 30 assets; collective insurance schemes, micro-credits, financial instruments for transferring risks, as 31 agricultural insurance and Payments for Ecosystem Services (PES) (Dávila, 2016; Hardoy and Velásquez, 32 2016; Lemos et al., 2016; Porras et al., 2016; Kalikoski et al., 2018). Small-scale household running 33 businesses in poor neighbourhoods develop adaptation strategies to keep business going, showing how 34 household level adaptation strategies are multipurpose (Stein et al., 2018; Stein, 2019). There are emerging 35 interinstitutional communities of practice with the purpose of sharing practices and lessons learned (ECLAC, 36 2013; ECLAC, 2015; ECLAC, 2019a). 37 38 There is also increasing evidence of human mobility associated with climate change and disaster risk (IOM, 39 2021) and the adoption of sustainable tourism, diversification of livelihoods strategies, climate forecasts, 40 appropriate construction techniques, neighbourhood layout, integral urban upgrading initiatives, territorial 41 and urban planning, regulatory frameworks, water harvesting and nature-based solutions (NbS) (Stein and 42 Moser, 2014; Hardoy and Mastrangelo, 2016; Almeida et al., 2018; Barbier and Hochard, 2018a; Desmaison 43 et al., 2018; Satterthwaite et al., 2018; Villafuerte et al., 2018; Hidalgo, 2020; Satterthwaite et al., 2020). 44 Mostly, socio-economical and socio-political factors which show safety and continuity measures are critical 45 enablers of adaptation. 46 47 At municipal level study in Central America highlighted that adaptive capacity in rural areas is associated 48 with basic needs satisfaction (safe drinking water, school, quality dwelling, gender parity index), access to 49 resources for innovation and action (road density, economically active population with non-agricultural 50 employment, and rural demographic dependency ratio), and access to credit and technical support 51 (Bouroncle et al., 2017). 52 53 CSA adaptation initiatives to reduce poverty, improve livelihoods and achieve sustainable development 54 range in scale and scope, from planned and collective interventions to autonomous and individual actions. 55 Many of them are bottom up, community-led initiatives together with civil society organizations; others are 56 government-led, including local governments, or a combination of them (McNamara and Buggy, 2017; 57 Berrang-Ford et al., 2021). Vulnerable groups are a focus to achieve equity at planning and as a target Do Not Cite, Quote or Distribute 12-81 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 including mainly rural low-income, Indigenous Peoples and women and migrants in most references. 2 Responses detected were focused on behavioural and cultural, followed by ecosystem-based responses, 3 institutional, and technological/infrastructural responses. Out of 55 articles analysed from CSA (Berrang- 4 Ford et al., 2021) about poverty, equity and adaptation options, half of them covered adaptation planning and 5 early implementation but only 2% could show evidence of risk reduction associated with adaptation efforts. 6 7 Tensions and conflicts may result from differing perceptions and knowledge on vulnerabilities and risk 8 which can hinder the acceptance of adaptation measures and implementing stronger adaptive or preventive 9 actions (Miranda Sara et al., 2016). There is a need to better understand complex interactions and community 10 responses to climate change in the Amazonian and Andean region. Climate change hotspot impacts, showed 11 that poverty reduction measures alone were not enough to improve adaptive capacity, as people will not 12 necessarily invest to enhance them (Pinho et al., 2014; Filho et al., 2016; Nelson et al., 2016; Lapola et al., 13 2018; Zavaleta et al., 2018). Current adaptation strategies and methods may be neglecting cultural values, 14 even eroding them, in Peruvian Andes, pointing that success of adaptation practices is tied to deep cultural 15 values (Walshe and Argumedo, 2016). 16 17 Limits to adaptation include access to land, territory and resources (Mesclier et al., 2015), poor labour 18 opportunities coupled with knowledge gaps, weak multi actor coordination, and lack of effective policies and 19 supportive frameworks (Berrang-Ford et al., 2021). 20 21 Low participation of women in income earning opportunities contrasts with their role in unpaid activities 22 (ECLAC, 2019b). Despite progresses, gender differences in labour markets remain an unjustifiable form of 23 inequality (OIT, 2019) and women easily fall back to the informal labour market during crisis situations such 24 as those generated by climate events (Collodi et al., 2020). 25 26 Participatory processes are leveraging adaptation measures throughout CSA; they contribute to prioritization 27 of specific adaptation measures as well as strengthening local capacities. Showing that climate adaptation 28 needs to be part of larger transformation processes to reduce vulnerability drivers (Stein and Moser, 2015; 29 Stein et al., 2018; Stein, 2019) but stronger national policies interlinking poverty and inequality reduction to 30 adaptation, considering the coupled human-environmental systems to comprehend poor and vulnerable 31 groups' capacity to adapt are urgent. CSA does not fare very well, and several downward trends might 32 become even more acute. More effective decisive actions need to be undertaken coupled with inclusive long- 33 term planning to protect the poor and improve their underlying conditions, to meet the SDG. 34 35 12.5.8 Cross-cutting Issues in the Human Dimension 36 37 12.5.8.1 Public policies, social movements and participation 38 39 Public policies related to adaptation must be seen in the wider context of environmental policies and 40 governance, as they usually address climatic processes in synergy with other environmental and 41 socioeconomic drivers (very high confidence) (Ding et al., 2017; Aldunce Ide et al., 2020; Comisión 42 Europea, 2020; Lampis et al., 2020; Scoville-Simonds et al., 2020). Some people rather point to education, 43 sanitation or social assistance, among other sectors (Bonatti et al., 2019). In Brazil, for example, it would be 44 difficult to clearly separate climate change adaptation and urban policies (high confidence) (PBMC, 2016; 45 Barbi and da Costa Ferreira, 2017; Marques Di Giulio et al., 2017; Empresa de Pesquisa Energética, 2018; 46 Checco and Caldas, 2019; Canil et al., 2020). 47 48 Many public policies related to climate change have become symbolic, in conflict with prevailing economic 49 policies and practices (medium confidence: low evidence, high agreement). Urban adaptation plans can be in 50 conflict with other policies and there may exist insufficient support in multiple areas such as social attitudes 51 and behaviour, knowledge, education and human capital, finance, governance, institutions and policy 52 (Villamizar et al., 2017; Koch, 2018). Some policies around climatic related displacements and migrants 53 have been considered in NDCs (Priotto and Salvador Aruj, 2017; Yamamoto et al., 2018; de Salles Cavedon- 54 Capdeville et al., 2020). 55 56 As there are asymmetries among populations regarding the vulnerability and benefits of adaptation, along the 57 lines of gender, age, socioeconomic conditions and ethnicity, it has been noticed that adaptation policies and Do Not Cite, Quote or Distribute 12-82 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 programs must be adequate to diverse conditions and actors (very high confidence) (Kaijser and Kronsell, 2 2014; Walshe and Argumedo, 2016; Baucom and Omelsky, 2017; Harvey et al., 2018). 3 4 Effective adaptation and mitigation depend on policies and measures at multiple scales, especially on the 5 involvement of the more exposed and vulnerable people. The participation of experts, communities and 6 citizens has shown to be effective (FAO and Fundación Futuro Latinoamericano, 2019) particularly through 7 partnership of grassroots organizations with impoverished communities providing valued expertise and 8 capacities to support the implementation of government climate resilience strategies (World Bank Group, 9 2015). More inclusive planning processes correspond to higher climate equity and justice outcomes in the 10 short term, but also an emphasis on building dedicated multi-sector governance institutions may enhance 11 long-term programs stability, while ensuring civil society voice in adaptation planning and implementation 12 (Chu et al., 2016). Some local organizations and people have succeeded when they were in charge of their 13 own resiliency efforts, where international projects and protocols proved less effective (Doughty, 2016). At 14 times, decentralized governmental programs have tried to increase public responsiveness to the adaptation 15 needs of people; however, proving to only be mildly successful and provoking the mobilization of 16 communities against existing governance structures (Thompson, 2016). 17 18 Indigenous knowledge and local knowledge (IK and LK) participation is thought to be more considered in 19 adaptation policies, as it has good results (high confidence) (Nagy et al., 2014b; Jurt et al., 2015; Arias et al., 20 2016; Stensrud, 2016). IK has been adaptive for long periods in the Andes (Cuvi, 2018), but there might be 21 limits to adaptation in the face of present climatic and other environmental and socioeconomic drivers 22 (Postigo, 2019). Approaches integrating IK with more formal sciences, to address research and policies, have 23 improved adaptation processes, but they are no exempt of complications (high confidence) (Doswald et al., 24 2014; Metternicht et al., 2014; Tengö et al., 2014; Drenkhan et al., 2015; Keenan, 2015; Lasage et al., 2015; 25 Camacho Guerreiro et al., 2016; Hurlbert and Gupta, 2016; Roco et al., 2016; Santos et al., 2016; Walshe 26 and Argumedo, 2016; Uribe Rivera et al., 2017; Kasecker et al., 2018; Cuesta et al., 2019; Ulloa, 2019; 27 Ariza-Montobbio and Cuvi, 2020). More interdisciplinary and transdisciplinary research helps to better 28 understand and manage the relationship between governance, implementation, management priorities, wealth 29 distribution and trade-offs between adaptation, mitigation and the Sustainable Development Goals (SDG). 30 31 Representations of climate change can also emerge as critiques and resistances, that expose that climate 32 change labelled politics or interventions have posed even bigger risks, or do not address poverty issues 33 (medium confidence: medium evidence, high agreement) (Lampis, 2013; Pokorny et al., 2013; Ojeda, 2014). 34 Indigenous and social movements have joined with climate justice activists, claiming for action against 35 climate change (Hicks and Fabricant, 2016; Ruiz-Mallén et al., 2017; Charles, 2021). The Bolivian Platform 36 against Climate Change, a coalition of civil society and social movement organizations working to address 37 the effects of global warming in Bolivia and to influence the broader global community, reflects an 38 innovative dimension that, albeit at time conflictual, has flagged how increasing climate variability hinders 39 the right of Indigenous Peoples to the conservation of their culture and practices and illustrates how grass- 40 root movements are increasingly appropriating climate change policy in the region (Hicks and Fabricant, 41 2016). Social movements have engaged with international networks as Blokadia, which surged after COP 23, 42 whose vindications try to go beyond the protection of the environment, delving into issues of democracy and 43 resource control (Martínez-Alier et al., 2018). 44 45 Many social movements address adaptation to climate change. Some engage and participate in policy and 46 planning, often having good results at the local level. On the contrary, top-down approaches without 47 participation have shown to be less effective (high confidence) (Krellenberg and Katrin, 2014; Nagy et al., 48 2014b; Stein and Moser, 2014; Ruiz-Mallén et al., 2015; Sherman et al., 2015; Waylen et al., 2015; Bizikova 49 et al., 2016; Chelleri et al., 2016; Merlinsky, 2016; Villamizar et al., 2017). 50 51 Some conflicts in which the direct biophysical impacts of climate change play a major role can unleash 52 social protests and strengthen social movements (Section 12.6.4). In Cartagena, since 2010, the increase in 53 precipitation increasingly impacted the barrio Policarpa, causing the residents to claim solutions for the 54 problems caused by the coupled effect of flooding and industrial pollution. Also, in El Cambray II, in 55 Guatemala City, in 2015 the nearby hill collapsed, causing the death of 280 people, 70 disappeared and the 56 destruction of hundreds of homes. The affected community entered into a conflict with the municipality 57 asking for resettlement and a reform of land-use planning (Stein Heinemann, 2018). Do Not Cite, Quote or Distribute 12-83 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 12.5.8.2 Perceptions 3 4 Perception and understanding of climate change can be seen as an adaptive feature. In CSA, the 5 consciousness of it as a threat is burgeoning, a situation related to a growth in climate justice activism, as 6 well as to the occurrence of extreme weather events of all kinds (high confidence) (Forero et al., 2014; 7 Magrin et al., 2014; Capstick et al., 2015). Perception is positively associated across countries with the 8 Human Development Index and ND-Gain Readiness Index, and negatively associated with the Vulnerability 9 Index, and within countries, with the education level, while they are negatively associated with the degree of 10 political affinity for the market economy (Azócar et al., 2021). Anyhow, some communities do not associate 11 their problems with the scientific concept, so discussions as if it is human induced, the causes, or relations 12 with other problems, can become irrelevant (Sapiains Arrué and Ugarte Caviedes, 2017). Even communities 13 affected by the same changes do not necessarily perceive them in the same way (Bonatti et al., 2016). The 14 interpretations of change, its causes and effects, can widely vary (Paerregaard, 2018; Scoville-Simonds, 15 2018). Rather than adapting to climate change, some peoples adapt climate change to their social worlds 16 (Rasmussen, 2016a). 17 18 Perceptions tend to be different in rural and urban areas (Sherman et al., 2015). In the rural areas, it is highly 19 related with temperature rise and changes in rainfall patterns, changes in agriculture (pests, calendars), 20 biodiversity loss, solar radiation or changes in the oceans, and their impacts sometimes are related or even 21 more attributed to socioeconomic and environmental drivers, and also related with financial negative 22 outcomes (high confidence) (Infante and Infante, 2013; Postigo, 2014; Jacobi et al., 2015; Barrucand et al., 23 2017; Harvey et al., 2018; Martins and Gasalla, 2018; Meldrum et al., 2018; Córdoba Vargas et al., 2019; 24 Leroy, 2019; Viguera et al., 2019; Gutierrez et al., 2020; Iniguez-Gallardo et al., 2020; Lambert and Eise, 25 2020). In places as the Amazonia, there is an increased perception with age (Funatsu et al., 2019). In 26 Mediterranean Chile, younger, more educated producers and those who own their land tend to have a clearer 27 perception than older, less educated, or tenant farmers, but they do not have a clear perception or how it may 28 affect their yields and farming operation (Roco et al., 2015). In some dry and humid Ecuadorian montane 29 forests, peasants perceive in the same way as scientific data, but they are at odds to predict the changes and 30 consider that they may not be prepared and only can be reactive (Herrador-Valencia and Paredes, 2016). In 31 an Andean community, perceptions of climate change are homogeneous and do not vary according to 32 gender, age or ethnicity (Cáceres-Arteaga et al., 2020). Among representatives of five municipalities of 33 Lima, it was found that climate change is not well understood and they have trouble distinguishing it from 34 other environmental issues (Siña et al., 2016). In an Amazonian region, farmers provided a more accurate 35 description than regional institutions of how it affects the local livelihood system (Altea, 2020). In Cuenca 36 Auqui peasants attribute recently experienced challenges in agricultural production mainly to perceived 37 changes in precipitation patterns, but statistical analyses of daily precipitation records at nearby stations do 38 not corroborate those perceived changes (Gurgiser et al., 2016). 39 40 12.5.8.3 Gender and intersectionality 41 42 There is ample empirical evidence that the impacts of climate change are not of equal scope for men and 43 women. Women, particularly the poorest, are more vulnerable and are impacted in greater proportion. Often, 44 for several economic and social reasons, they have less capacity to adapt, further widening structural gender 45 gaps (high confidence) (Box 7.4; Arana Zegarra, 2017; Casas Varez, 2017; Segnestam, 2017; Acosta et al., 46 2019; Aldunce Ide et al., 2020; Olivera et al., 2021; Silva Rodríguez de San Miguel et al., 2021). Gender 47 equity is thought to be central to discussions on climate change adaptation policies. In issues such as 48 drinking water, energy, natural disasters, impacts on health and agriculture, capacity to migrate, women 49 (poor women in particular) are affected in greater proportion, further widening structural gender gaps. In a 50 rural community vulnerable to drought, short-term coping was more common among the women, especially 51 among female heads of household, while adaptive actions were more usual among the men; there are 52 gendered inequalities in access to and control over different forms of capital that lead to a gender- 53 differentiated capacity to adapt, where men are better able to adapt and women experience a downward 54 spiral in their capacity to adapt and increasing vulnerability to drought (Segnestam, 2017). 55 56 However, women are not always the more vulnerable group. While in a broad sense climate change impacts 57 more severely on women, there are situations where they have reacted, adapted better to, and been more Do Not Cite, Quote or Distribute 12-84 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 resilient. Grassroots women self-help groups can be active agents of change for their communities, designing 2 and delivering gender-responsive adaptation solutions (Huairou Commission, 2019). Some studies suggest 3 that women establish a friendlier relationship with the environment and towards natural resources; studies on 4 masculinities and environment confirm this tendency (Brough et al., 2016). In a multi country study, some 5 female headed households tend to be slightly less vulnerable and more resilient than male headed 6 households, even though some exceptions were found when looking at sub-groups (Andersen et al., 2017). In 7 Chile, women are more likely to modernize irrigation and infrastructure, and gender appears as an important 8 element for drought adaptation (Roco et al., 2016). A change to agro-ecological practices has improved 9 gender equalities and adaptive capacity to climate change (Cáceres-Arteaga et al., 2020). 10 11 Recent studies emphasize that a gender approach to social inequalities ought to move beyond just looking at 12 men and women as experiencing the impacts in a differentiated manner; rather, an intersectional analysis 13 illuminates how different individuals and groups relate differently to climate change, due to their 14 situatedness in power structures based on context-specific and dynamic social categorizations (high 15 confidence) (Kaijser and Kronsell, 2014; Djoudi et al., 2016; Thompson-Hall et al., 2016; Olivera et al., 16 2021). Thus, the relationship between gender and adaptation demands an analytical framework that connects 17 environmental problems with social inequalities in a complex way (Godfrey, 2012). An intersectional 18 approach contributes to better capture the diversity of adaptive strategies that men and women adopt vis-à- 19 vis climate change. Particular constellations of race, gender, class, age or nationality reveal more complex 20 realities (high confidence). 21 22 12.5.8.4 Migrations and displacements 23 24 Migration and displacements are multi-causal phenomena, and climate may exacerbate political, social, 25 economic or other environmental drivers (high confidence) (Kaenzig and Piguet, 2014; Brandt et al., 2016; 26 Priotto and Salvador Aruj, 2017; Sudmeier-Rieux et al., 2017; Radel et al., 2018; Heslin et al., 2019; 27 Hoffmann et al., 2020; Silva Rodríguez de San Miguel et al., 2021). In the region there are many case 28 studies, but data to assess and monitor precisely the effects of climate -and weather- related disasters in 29 migration and displacements in a broad perspective is still inaccurate (Priotto and Salvador Aruj, 2017; 30 Abeldaño Zuñiga and Fanta Garrido, 2020). The most common climatic drivers include tropical storms and 31 hurricanes, heavy rains, floods and droughts (Kaenzig and Piguet, 2014). Positive climatic conditions also 32 can facilitate migration (Gray and Bilsborrow, 2013). Peru, Colombia and Guatemala are amongst the 33 countries with the largest average displacements caused by hydro meteorological causes; Brazil had 295,000 34 people displaced because of disasters in 2019 (Global Internal Displacement Database, https://www.internal- 35 displacement.org/database/displacement-data). 36 37 These processes can be interpreted as impacts in vulnerable peoples, but also as adaptation strategies to 38 manage the risks and reduce the exposure, when people continue with their lives, temporary or permanently, 39 in a different but stable situation, or when members of the families send remittances to those that remain in 40 the affected areas (Section 7.4.3.2; Cross-Chapter Box MIGRATE in Chapter 7). The remittances create 41 opportunities for adaptive capacity building, as they reduce some vulnerabilities in the form of 42 infrastructures, agricultural supplies, food, education or health, as in northern CA (NU CEPAL, 2018). 43 Anyhow, migration as adaptation is not available to everyone (Kaenzig and Piguet, 2014), and the idea has 44 also been contested as it may not help to overcome structural problems or point to in situ options (Radel et 45 al., 2018; Ruiz-de-Oña et al., 2019). The causal processes are complex. Surveys of migrants usually find that 46 the main reported reason for migration is to find a job or to increase the household income (Wrathall and 47 Suckall, 2016; OIM, 2017; Radel et al., 2018), but the underlying reason for the lack of job or income is 48 rarely examined, and at times may be related with climatic hazards. 49 50 Migration most often originates in rural areas, with people moving to other rural or urban areas within their 51 home countries (Table Cross-Chapter Box MIGRATE 1 in Chapter 7). In the Amazon, approximately 80% 52 of the population is concentrated in cities due to rural-urban migrations in search of better income, 53 livelihoods and services, in cases associated with extreme floods and droughts (Pinho et al., 2015). In 54 Ecuador, environmental variables are most likely to enhance international than internal migration (Gray and 55 Bilsborrow, 2013). Hurricanes have been seen as positive triggers for international migration in CA (Spencer 56 and Urquhart, 2018). In the highlands of Peru, there are different patterns, including daily circular migration Do Not Cite, Quote or Distribute 12-85 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 to combine the scarce income from agricultural production with urban income, rather than abandoning the 2 farming land (Milan and Ho, 2014; Zimmerer, 2014; Bergmann et al., 2021). 3 4 Migration to cities can mean opportunities for migrants and for the urban areas, but also can worsen the 5 problems, as urban poor people can become even more exposed and vulnerable, and the pressure on urban 6 capacities may not be well absorbed (high confidence) (Chisari and Miller, 2016; Gemenne et al., 2020). 7 Internal migration to cities is likely to exacerbate pre-existing vulnerabilities related to inequality, poverty, 8 indigence and informality (Warn and Adamo, 2014). Immigration can make cities/residents more vulnerable 9 to climate change risks (Section 12.5.5; Section 12.5.7). Groups as children, Indigenous Peoples or the poor 10 are usually amongst the most vulnerable in the migrations and displacements, which poses challenges to 11 national policies and international aid (Sedeh, 2014; Gamez, 2016; Ulla, 2016; Priotto and Salvador Aruj, 12 2017; Ramos and de Salles Cavedon-Capdeville, 2017; Amar-Amar et al., 2019; Gemenne et al., 2020). In 13 forced migration or displacement by climatic effects, women are prone to lose their leadership, autonomy 14 and voice, especially in new organizational structures imposed by authorities. This is especially the case in 15 temporary accommodation camps created after disasters, exacerbating differentiated vulnerabilities existing 16 (Aldunce Ide et al., 2020). International migration has become more dangerous and difficult as border 17 controls have become stricter, but programs such as the one of temporary agricultural workers from 18 Guatemala to Canada have proven to be successful (Gabriel and Macdonald, 2018). At the same time, 19 emigration may lead to the loss of IK and LK for adaptation (Moreno et al., 2020b). 20 21 Some areas are more sensitive to generate climatic migration: the Andes, the dry areas of the Amazonia, 22 northern Brazil, and the northern countries in CA (high confidence). Northeast Brazil would lose population 23 that will move to the south, deepening the existing inequalities (Oliveira and Pereda, 2020). In a study of 8 24 countries around the world, including Guatemala and Peru, a link was found between rainfall variability and 25 food insecurity which could lead to migration in areas of high prevalence of rainfed agriculture and low 26 diversification (Warner and Afifi, 2014). In CA, younger individuals are more likely to migrate in response 27 to hurricanes and especially to droughts (Baez et al., 2017). 28 29 The perception of gradual changes lowers the likelihood for internal migration, while sudden-onset events 30 increase movement (Koubi et al., 2016). On the other hand, it has been seen that extreme events like floods 31 or droughts can hinder population mobility, immobilizing them in their localities (Thiede et al., 2016). These 32 immobilized populations are supposed to face a double set of risks: they are unable to move away from 33 environmental threats, and their lack of capital makes them especially vulnerable to environmental changes 34 (Black et al., 2011). In CSA, migrating to the U.S. is becoming dangerous and expensive, as that country is 35 restricting the entries; these trends expose local populations to the risk of becoming immobile in the near 36 future in a place where they are extremely vulnerable (Ruano and Milan, 2014; McLeman, 2019). A survey 37 in Guatemala found no correlation between migration to the U.S. and severe food insecurity in households, 38 but the correlation became significant if the level of food insecurity was moderate, suggesting that families 39 in extreme hardship did not have the resources to migrate (Aguilar et al., 2019). At the same time, some 40 populations just have chosen not to move, as in Peru, where immobility in dissatisfied people is more likely 41 to be caused by attachment to place than resource constraints (Adams, 2016; Correia and Ojima, 2017). 42 Some populations have chosen to adapt relying in their IK and LK (Boillat and Berkes, 2013). 43 44 Migration is often the last resort for rural communities facing water stress problems (Magrin et al., 2014; 45 Ruano and Milan, 2014). In Bolivia, glacial retreat has not triggered new migration flows and had a limited 46 impact on the existing migratory patterns (Kaenzig, 2015). In SA, climatic variability increases the 47 likelihood of inter-province migration, rather than trapping populations. In a study of interprovincial 48 migration motivated by temperature, an exception arose in Bolivia, and even if that could suggest an 49 immobilized population (Thiede et al., 2016), it is not clear if they want to stay and adapt. In some cases, 50 people want to move but wait for relocation after the climate related disasters (Priotto and Salvador Aruj, 51 2017). 52 53 12.5.8.5 Financing 54 55 Climate change financing is unequally distributed among CSA countries (high confidence). Financing of 56 climate change adaptation remains very much delegated to multilateral and bilateral cooperation and the 57 governments in the region have heavily relied on it. Still, there are some concerns regarding justice in the Do Not Cite, Quote or Distribute 12-86 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 distribution of these funds (Khan et al., 2020). The UNFCCC has created financing mechanisms throughout 2 its functioning years, but there is a wide range of issues that can present challenges for access by the 3 recipients (Hickmann et al., 2019). These include; lack of technical capacity; difficulties in following the 4 procedures established by the various financial entities; and low levels of awareness about the need for 5 action, as well as the different sources of funds available. The fiscal policies of the different countries have 6 contributed to government financing in the fight against climate change (World Bank, 2021). Since the Paris 7 Agreement, countries have pledged NDCs which introduce the need to design and implement carbon budgets 8 with respective consideration of the efficiency and costs and benefits involved in each mitigation or 9 adaptation to climate change projects (Fragkos, 2020). 10 11 According to UNFCCC, Latin America and the Caribbean, for the period 2015­2016 obtained 22% of 12 climate finance from multilateral climate funds. In this section we use data from: 13 https://climatefundsupdate.org/data-dashboard, most of the reported information for Latin-American and the 14 Caribbean includes Mexico, since the scope of this chapter does not includes Mexico we have rely in the raw 15 data included in the data-dashboard mentioned in the link (see also: Guzmán et al. (2016)). 76% went to 16 mitigation projects with the remaining 24% going to adaptation. Of the total finance provided by the 17 multilateral climate funds to the Region, 51% took the form of concessional loans, while 47% was provided 18 as grants. For the region, approvals in the 2015­2016 period were concentrated in Argentina, Chile, Brazil, 19 and Colombia, where large-scale mitigation projects were launched supported by the Green Climate Fund 20 (GCF) and the Clean Technology Fund (CTF). For the period 2003-2019, total contribution to South 21 America and the Caribbean is about USD 3,558 million. The largest contributors to climate finance in the 22 region come from the GCF, which approved USD 824.2 million for 23 projects. Brazil is the top recipient 23 with USD 195 million, followed by Argentina with about USD 162 million. The second provider is the 24 Amazon Fund with USD 717 million assigned to 102 projects in Brazil. In 2018, the CTF has become the 25 third source of financing with USD 483 million dollars approved for 24 projects; the main recipient is Chile 26 with USD 16,207 million followed by Colombia with USD 170 million. The five largest projects approved in 27 the region in 2018 were through the GCF. Brazil (USD 195 million) received support for reducing energy 28 intensity across Brazilian cities, while Argentina (USD 103 million) received support to scale up investments 29 by Small and Medium sized Enterprises (SMEs) in renewable energy and energy efficiency. In both cases 30 finance is predominantly provided as concessional loans. 31 32 Climate financing in CSA is mainly focused on mitigation actions (high confidence). In South America and 33 the Caribbean, 73% (USD 2,579 million) of funding to date has supported mitigation. Only 21% (USD 761 34 million) of the funding supports adaptation projects and the remaining 4% (USD 217 million) supports 35 multi-focus projects. Of the 51 new projects in South America and the Caribbean approved in 2018-2019, the 36 GCF financed USD 508 million in ten projects. Amazon Fund was next with USD 81 million in 10 projects. 37 While 32 the GCF focuses on large and transformative projects and programs and on a broader reform of the 38 policy framework in the Region, the Amazon Fund targets smaller project interventions. 39 40 Climate finance in the region is concentrated in Brazil receiving one third of the region's funding, and 41 41 mitigation activities receiving more than six times that of adaptation from multilateral climate funds. By the 42 size of its PGB, Brazil is receiving the largest amount of financing; this leaves the poorest countries with 43 little or no financing and therefore reinforces a vicious circle of poverty and vulnerability. If this is due to 44 Brazil being more successful presenting eligible projects, lack of commitment from other developing 45 countries or some other structural factors is an open question. In any case, compensation schemes for the 46 most vulnerable countries appear as required, given the differences in vulnerability to climate damages 47 (Antimiani et al., 2017). This is aggravated by the fact that funds management is in the hands of 48 supranational entities while inequalities remain in regions within a country, particularly in countries highly 49 centralized as is the case for countries in the region. 50 51 COVID-19 recovery plans can present synergistic effects for climate change adaptation (medium confidence: 52 low evidence, high agreement). A key decision point for adaptation will be how the world responds to the 53 pandemic. The global recovery can serve as a catalyst to increased and more equitable climate financing. 54 Globally, recovery packages will likely have the power to change the global trajectory towards meeting the 55 targets of the Paris Agreement and building a more just future (Forster et al., 2020). Several factors are 56 relevant to the design of economic recovery packages: the long run economic multiplier, contributions to the 57 productive asset base and national wealth, speed of implementation, affordability, simplicity, impact on Do Not Cite, Quote or Distribute 12-87 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 inequality, and various political considerations (Hepburn et al., 2020). A key objective of any recovery 2 package is to stabilize expectations, restore confidence, and to channel surplus desired savings into 3 productive investment. However, `business as usual' implies temperature increases over 3°C, implying great 4 future uncertainty, instability, and climate damages. An alternative way to restore confidence is to steer 5 investment towards a productive and balanced portfolio of sustainable physical capital, human capital, social 6 capital, intangible capital, and natural capital assets (Zenghelis et al., 2020), consistent with global goals on 7 climate change. Finally, any recovery package, including climate-friendly recovery, is unlikely to be 8 implemented unless it also addresses existing societal and political concerns--such as poverty alleviation, 9 inequality, and social inclusion--which vary from country to country. 10 11 12.5.9 Adaptation Options to Address Key Risks in CSA 12 13 This section integrates, in the table 12.10 below, the sectoral assessment of adaptation options (see Sections 14 12.5.1 to 12.5.8) with the eight key risks assessed in the region (see Section 12.4). Table 12.10 presents a list 15 of the summarized adaptation options, which are detailed in their adaptation sections, from 12.5.1 to 12.5.8 16 in this chapter. 17 18 19 Table 12.10: Adaptation options addressing key risks organized by sector. See the note at the end for descriptions of 20 the sector names abbreviations. 1. Risk of food insecurity due to frequent/extreme droughts T&F. ecosystems Ecosystem-based adaptation (EbA): Agroecosystem resilience practices O&C ecosystems Water Not Assessed (NA) Food Water infrastructure and irrigation; Nature-based solution (NbS) & Payment for ecosystem services (PES); Participatory water management; Multi-purpose water use Climate information services; Early warning system (EWS); Insurance; Land use planning; Low-Carbon Agriculture (LCA) strategies; Agroforestry; Indigenous Knowledge and Local knowledge (IK and LK) Cities NA Health and wellbeing EWS; Insurance; Participatory water management; Water infrastructure and irrigation Poverty and SD Community-based adaptation (CbA); Government and institutional support Human Dimension Participatory management; Incorporation of IK and LK in water and crop management; Education and communication 2. Risk to life and infrastructure due to floods and landslides T&F ecosystems NA O&C ecosystems NA Water NbS; Land-use regulation; EWS; Integrated risk management. Food NA Do Not Cite, Quote or Distribute 12-88 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report Cities Urban planning; Climate-adapted parameters in land use and building regulation; Intersectoral and multilevel governance; Slum upgrading; Social housing improvement; Urban control systems; CbA; Risk management plans; Integrated watershed management; Flood control programs; Environment protected areas; Households relocation; EWS; NbS; Mapping tools; Green-grey infrastructure (GGI); Water storage solutions; Wetland restoration; sustainable urban drainage systems (SUDS); low-impact development (LID); River restoration; Multifunctional landscapes; Improving basic sanitation services Health and wellbeing EWS; GGI; Community led and managed relocation; Insurance Poverty and SD Secure location; Social housing policies; EWS Human dimensions Education and communication 3. Risk of water insecurity T&F ecosystems Monitoring Systems; EbA; Forest protection and restoration; Watershed protection O&C ecosystems CbA; Land use and development regulation Water Water infrastructure and irrigation; NbS & PES; Participatory water management; Food Multi-purpose water use Management and planning; NbS; Soil and water conservation Cities Intersectoral and multilevel governance; CbA; Risk management plans; Integrated Health and wellbeing watershed management; Environment protected areas; NbS; GGI; Wetland restoration; Poverty and SD Improving basic sanitation services; Reservoir system Protection and restoration; National Adaptation Plans; Participatory water management NbS: Water harvesting; Equitable water distribution Human dimensions Participatory management; Incorporation of IK and LK in water management; Education and communication 4. Risk of severe health effects due to increasing epidemics T&F ecosystems NA O&C ecosystems NA Water Food Water infrastructure; Sanitation improvement Cities NA Health and wellbeing NA EWS; Health-climate surveillance systems; National plans on health; Communal Poverty and SD management; GGI; Protection and Restoration. CbA; Transparent democratic governance; Equitable services; Education Human dimensions Education and communication Do Not Cite, Quote or Distribute 12-89 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 5. Systemic risks of surpassing infrastructure and public service systems T&F ecosystems NA O&C ecosystems EWS; EbA; Territorial planning; CbA; Land use and development regulation; GGI Water Water infrastructure; Land-use regulation; Water retention capacity; EWS; Capacity Food building Cities NA Health and wellbeing Poverty and SD Urban planning; Climate-adapted parameters in land use and building regulation; Intersectoral and multilevel governance; Slum upgrading; Social housing improvement; CbA; Improving basic sanitation services; Micro wastewater treatment plants EWS; Vulnerability and risk maps; National Adaptation Plans; GGI Transparent, democratic governance Human dimensions NA 6. Risk of large-scale changes and biome shifts in the Amazon T&F ecosystems Monitoring Systems; EbA; Protected areas; Forest protection and restoration and O&C ecosystems restoration; Watershed protection NA Water Integrated water resource management Food Territorial planning Cities NA Health and wellbeing Protection and restoration Poverty and SD Insurance; Micro-credits; PES; CbA Human dimensions Participatory management; Incorporation of IK and LK in forest management; Education and communication 7. Risk to coral reef ecosystems due to coral bleaching T&F ecosystems NA O&C ecosystems Zoning schemes; MPAs; EbA; CbA; Adhesion of international treaties Water NA Food NA Do Not Cite, Quote or Distribute 12-90 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report Cities NA Health and wellbeing Protection and restoration Poverty and SD NA Human dimensions NA 8. Risks to coastal socio-ecological systems due to sea level rise, storm surges and coastal erosion T&F ecosystems NA O&C ecosystems EbA; Planned relocation; GGI Water NA Food NA Cities Urban planning; Climate-adapted patterns in land use and building regulation; Health and wellbeing Intersectoral and multilevel governance; CbA; Risk management plans; Households relocation; NbS; GGI GGI; Communal management; Protection and restoration Poverty and SD Secure location; CbA relocation Human dimensions Participatory management; Education and communication 1 Table Notes: 2 Some sectors are presented by abbreviations: Terrestrial and freshwater ecosystems and their services (T&F. 3 ecosystems); Ocean and coastal ecosystems and their services (O&C ecosystems); Food, fibre and other ecosystem 4 products (Food); Cities, settlements and key infrastructure (Cities); Poverty, livelihood and sustainable development 5 (Poverty and SD); Cross cutting issues in the Human Dimension (Human Dimensions). 6 7 Feasibility Assessment of Adaptation Options 8 12.5.10 9 10 This section assesses the feasibility of selected adaptations options by sector, relevant for CSA, in five 11 dimensions (economic, technological, institutional, social, environmental and geophysical), according to the 12 methodology developed by Singh et al. (2020a). Table 12.11 shows the summary of results and Table 13 SM12.7 the details of the assessment and the supporting literature. 14 15 16 Table 12.11: Feasibility assessment of selected adaptation options for CSA region. System Adaptation Evidence Agreement Dimension assessed option Economic Technologi Institutio Social Environme Geophy cal nal ntal sical Do Not Cite, Quote or Distribute 12-91 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report Food, Agroforest Medium High Insignific Mixed Significa Mixed Insignifica Mixed fibre ry nt barriers effect and ant effect nt effect other ecosyst barriers barriers em produc ts Health Early Robust High Insignific Mixed Significa Mixed Insignifica Mixed and warning nt barriers effect wellbei systems ant effect nt effect ng barriers barriers Water Multi-use Robust Medium Insignific Mixed Mixed Mixed Mixed Insigni of water effect effect effect ficant storage ant effect barrier approache s s barriers Freshw Ecosystem Medium High Mixed Mixed Insignifica Insignifica Insigni ater -based effect effect nt barriers nt barriers ficant and adaptation Insignific barrier terrestr (EbA) ant s ial barriers ecosyst ems 1 2 3 12.5.10.1 Food, fibre and other ecosystem products - Agroforestry 4 5 For the agri-food systems, the adoption of agroforestry provides a more diverse and sustainable agricultural 6 production, where farmers maintain or improve their current production by incorporating suitable trees that 7 ameliorate climatic conditions. Thus, in the same unit of land, these systems incorporate exotic tree species 8 or managed native forests into farming systems allowing the simultaneous production of trees, crops and 9 livestock with different spatial arrangements or temporal sequences. On the other hand, it is recognized that 10 the initial investment and time until trees start to produce may create economic vulnerability. Therefore, 11 there is a need to design adequate programs and allocate resources for agroforestry systems implementation, 12 as well technical assistance and training (medium confidence). Also, some market schemes such as payment 13 for ecosystem services and certification can assist to reduce this vulnerability. 14 15 12.5.10.2 Health and Wellbeing - Early Warning Systems 16 17 For the health sector, we assessed the barriers and facilitators for the implementation of climate-driven early 18 warning systems under natural disasters and epidemic situations. We found institutional dimensions as 19 potential barriers, including the legal and regulatory feasibility, the institutional capacity and administrative 20 feasibility, transparency, and political acceptability (high confidence). The fewest barriers were identified for 21 the economic and environmental dimensions. 22 23 One of the main institutional challenges is the lack of policy with climate-health linkages. Opportunities 24 include a national plan for the health sector to address the impacts of climate by formalizing collaborations 25 via agreements (MOUs). Another key barrier is that relatively few institutions in the region have the human 26 technical and administrative capacity to implement and operate an EWS. Regional platforms may provide a 27 solution for technical assistance at national levels. 28 29 On the other hand, the economic dimensions had relatively few barriers, although the initial costs of 30 designing, implementing, equipping, and maintaining the system are a potential barrier for health sectors 31 with reduced budgets. However, the health benefits and economic savings (due to averted epidemics or 32 damages from disasters) may offset these costs. The resilience built in the health sector by these systems may Do Not Cite, Quote or Distribute 12-92 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 be applicable to other economic sectors that can benefit from an early warning of an oncoming extreme 2 event and associated health impacts. 3 4 12.5.10.3 Water - Multi-use of water storage approaches 5 6 For the water sector, geophysical and economic dimensions do not pose a major barrier due to the potential 7 reduction of flood hazard exposure, physical-technical viability of project implementation, different suitable 8 economic mechanisms for joint public-private financing and more efficient water use. However, limited 9 institutional capacities and the social-environmental impacts of large water infrastructure (Section 12.5.3) 10 reduce the institutional, social, environmental and, to some extent, technological feasibility. This may be a 11 potential barrier to the adaptive approach of multi-use water storage (medium confidence). 12 13 12.5.10.4 Freshwater and terrestrial ecosystems - Ecosystem-based adaptation (EbA) 14 15 In the terrestrial and freshwater ecosystems sector, we assessed the feasibility of implementing EbA options 16 in the CSA region. Given that EbA encompasses a wide range of projects, techniques and political and 17 socioeconomic arrangements, extreme care should be taken to apply these general findings to particular 18 cases. EbA can enhance food sovereignty and carbon stocks and foster SDG by protecting and restoring 19 ecosystems health and productivity. EbA is a strategy that frequently involves bottom-up decision making 20 and local communities' empowerment and usually contributes to inequality reduction. EbA tends to benefit 21 vulnerable groups, but aspects such as the impact on socioeconomic inequalities when implemented should 22 be taken into account. 23 24 In general, EbA does not require high technologies for local communities. However, limitations in technical 25 assistance and funding for specific key technologies and training may act as a barrier for EbA adoption 26 (medium confidence). EbA practices can reduce risk in several ways by increasing awareness among 27 communities and providing food diversity and production. EbA is recognized as a desirable policy for most 28 stakeholders in CSA, particularly for being a strategy that incorporates environmental and social concerns. 29 Nonetheless, it is important that all stakeholders agree on the goals and methods for EbA to be effective. 30 Lack of institutional coordination, clear goals and strategies were identified as a potential barrier for EbA 31 implementation. EbA is heavily based in local and Indigenous knowledge, as well as ecological academic 32 knowledge. 33 34 For the adaptation options analysed, significant barriers and mixed effects were observed for the institutional 35 dimension, which indicates the relevance of the design and implementation of public policies and 36 institutional arrangements for effective adaptation in the region. Considering the results, there is a need to 37 advance initiatives, programs and projects that facilitate adaptation to climate change. In the same way, 38 barriers were evidenced in the technological dimension, which indicates the importance of increasing access 39 and diffusion of appropriate techniques and technologies in order to face the challenges of climate change in 40 the region. 41 42 43 12.6 Case Studies 44 45 12.6.1 Nature-based Solutions in Quito, Ecuador 46 47 Nature-based Solutions (NbS) are related to the maintenance, enhancement, and restoration of biodiversity 48 and ecosystems as a means to address multiple concerns simultaneously (Kabisch et al., 2016). NbS can 49 trigger sustainability transitions. For example, conservation and restoration of natural ecosystems are prone 50 to promote synergy between mitigation, adaptation and sustainable development. Ecosystem-based 51 Adaptation- EbA can be seen as a type of NbS deployed in response to climate change vulnerability and risk 52 (Greenwalt et al., 2018), combining the objectives of reducing the vulnerability of human and increasing the 53 resilience of natural systems (IPCC, 2014). 54 55 The Municipal Quito District in Ecuador covers 4235 km2 of mountainous territory that ranges from 500 to 56 5000 m.a.s.l. That territory has followed a pattern of urbanization common in Latin America: its population 57 has increased from around 500,000 people in the 1970s, to nearly 3 million inhabitants by 2020, of which Do Not Cite, Quote or Distribute 12-93 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 80% live in urban areas (Municipio del Distrito Metropolitano de Quito, 2016). A massive inflow of people 2 immigrated in the early 1970s due to various causes, including the search for the rents created as a result of 3 the oil boom in the Ecuadorian Amazon, better working conditions, health, education and cultural services, 4 in comparison with the rural areas or in mid-sized cities. As a result, the city underwent an exponential 5 growth, claiming valuable agricultural and forestry areas, and natural ecosystems, in the peripheries. Many 6 of the new neighbourhoods were established through land invasions or informal markets, in many cases over 7 steep slopes, in water sources and agricultural or conservation areas (high confidence) (Cuvi, 2015; Gómez 8 Salazar and Cuvi, 2016). That exponential population growth, coupled with urban sprawl, poses many 9 challenges to the city, including those related to climate change. 10 11 Mean air temperature and annual rainfall (measured through instruments since 1891 and inferred through 12 historical records of rogation ceremonies since 1600), are increasing, combined with an increase in 13 seasonality (i.e., longer periods of drought) and extreme weather events, particularly stronger precipitations 14 (Serrano Vincenti et al., 2017; Domínguez-Castro et al., 2018). Two impacts related to warmer air conditions 15 are the displacement of the freezing line currently placed at 5100 m.a.s.l. (Basantes-Serrano et al., 2016), 16 followed by glacier retreat and the upward displacement of mountainous ecosystems (very high confidence) 17 (Vuille et al., 2018; Cuesta et al., 2019). The key ecosystem that regulates water provision for the city is the 18 paramo, and only about 5% of this process is related with glaciers, so the combined effects of climate change 19 on both systems, coupled with land use change and fires, can reduce the availability of water for agriculture, 20 human consumption and hydropower. Other important climatic hazards and impacts are the increase of solar 21 radiation, the heat island effect and fires (high confidence) (Anderson et al., 2011; Armenteras et al., 2020; 22 Ranasinghe et al., 2021). Almost half of the days of each year, Quito's population is exposed to levels of UV 23 radiation above 11 according to the World Health Organization scale (Municipio del Distrito Metropolitano 24 de Quito, 2016). 25 26 Various policies, programs and projects have been created for the promotion of urban green spaces, 27 protected areas, water sources and watersheds monitoring, conservation and ecosystem restoration, air 28 pollution monitoring and control, and urban agriculture. Among those actions, three recent are commonly 29 highlighted. The first is the Fund for the Protection of Water (FONAG), established in 2000 with funds of 30 national and international organizations, to promote the protection of the water basins that supply most of the 31 drinking water. It is a PES-Scheme (Payment for Ecosystem Services) enabled through a public-private 32 escrow. The projects include conservation, ecological restoration, and environmental education for a new 33 culture of water, in a context opposed to the commodification of natural resources (Kauffman, 2014; Bremer 34 et al., 2016; Coronel T, 2019). FONAG was innovative in the use of trust funds in a voluntary, decentralized 35 mechanism and has inspired more than 21 other water funds in the region; nevertheless, its narrative of 36 success has also been said to over-simplify and misrepresent some complex interactions between 37 stakeholders as well as within communities and their land management practices (Joslin, 2019). 38 39 The second highlighted initiative is the project AGRUPAR (Participative Urban Agriculture), launched as a 40 public initiative in 2002 with international cooperation funds at the beginning. It was aimed to provide 41 assistance to poorer urban and peri-urban populations, to initiate and manage orchards as well as domestic 42 animals such as chickens and guinea pigs, dedicated for self-sustenance and commerce. AGRUPAR provides 43 and finances training, seeds and seedlings, greenhouses, certifications and marketing support, spaces where 44 farmers can sell directly their products to consumers. In 2016, AGRUPAR gave assistance to more than 4000 45 farmers managing orchards of various scales that combined produce, annually, more than 500 tonnes. The 46 program has direct impacts on nutrition, generation of work for women, production of healthy food, 47 reduction of runoff, recycling of organic waste, social cohesion, among others (very high confidence) 48 (Thomas, 2014; Cuvi, 2015; Rodríguez-Dueñas and Rivera, 2016; Clavijo Palacios and Cuvi, 2017). 49 50 A third initiative is the creation of a municipal system of protected areas, locally named Áreas de 51 Conservación y Uso Sustentable (ACUS). This system covers an area of 1320 km², nearly one third of the 52 Municipal Quito District. Half of this landscape (680 km2) is covered by montane forests and paramos 53 (Torres and Peralvo, 2019). These forests provide direct water, food and fibres for about 20,000 people, and 54 indirectly a rural landscape for a growing number of urban citizens and foreign tourists that practice 55 ecotourism and look for fresh and healthy food. During the last three decades, this area has witnessed a high 56 density of public and private conservation and restoration efforts that aim to regain ecological integrity and 57 improve human well-being in deforested and degraded landscapes (Mansourian, 2017; Zalles, 2018; Wiegant Do Not Cite, Quote or Distribute 12-94 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 et al., 2020). Quito's system of protected areas constitutes a primary strategy for fostering links between 2 urban and rural citizens as a means of understanding the ecological dependence of urban metropolises to 3 their surrounding natural landscapes. Along the same lines, these areas constitute a key element to increase 4 the adaptive capacity of rural livelihoods and contribute to mitigating climate change through landscape 5 restoration, sustainable production and forest conservation (high confidence). 6 7 Other NbS' actions have been the restoration of small basins, locally named quebradas, under different 8 schemes of management and participation (medium evidence, very high agreement) (da Cruz e Sousa and 9 Ríos-Touma, 2018), or the transformation since 2013 of a larger portion of the old Quito airport into an 10 urban park. Nevertheless, Quito city still has to deal with challenges in social, economic, infrastructural and 11 environmental spheres. A major pending environmental issue is air pollution, as there is a high level of 12 pollutants affecting the city in general, and specially the most vulnerable groups (high confidence) 13 (Zalakeviciute et al., 2018; Alvarez-Mendoza et al., 2019; Estrella et al., 2019; Hernandez et al., 2019; 14 Rodríguez-Guerra and Cuvi, 2019). Another major issue is the continuous sprawl of new neighbourhoods, 15 mainly through informal processes, that diminish the urban resilience because of the destruction of 16 conservation and food production areas, sources of water, and the dispersion of settlements without primary 17 services, among other consequences (Gómez Salazar and Cuvi, 2016). 18 19 12.6.2 Anthropogenic Soils, an Option for Mitigation and Adaptation to Climate Change in Central and 20 South America. Learning from the "Terras Pretas de Índio" in the Amazon 21 22 Amazon Dark Earths (ADEs), also known as "Terras Pretas de Índio", are anthropogenic soils derived from 23 the activities associated to settlements and agricultural practices of pre-Hispanic societies in the Amazon 24 (Woods and McCann, 1999; Lehmann et al., 2003; Sombroek et al., 2003). Most of the ADEs identified so 25 far are 500 to 2500 years old (de Souza et al., 2019). According to Maezumi et al. (2018a) polyculture 26 agroforestry allowed the development of complex societies in the eastern Amazon around 4500 years ago. 27 Agroforestry was combined with the cultivation of multiple crops and the active and progressive increase in 28 the proportion of edible plant species in the forest, along with hunting and fishing. The formation of ADEs, 29 as a result of these activities, provided the basis for a food production system that supported a growing 30 human populations in the area (Maezumi et al., 2018a). 31 32 Amazon Dark Earths are the result of the accumulation and incomplete combustion of waste materials such 33 as ceramic artefacts and organic residues from harvest, weeding, food processing (including cooking) and 34 other activities (Lima et al., 2002; Hecht, 2003; Kämpf et al., 2003). ADEs are characterized by their 35 increased fertility in relation to adjacent soils; with high contents of organic carbon (C) (mainly as charcoal) 36 as well as inorganic nutrients, especially phosphorus (P) and calcium (Ca); and high Carbon/Nitrogen ratios 37 (high confidence) (Moline and Coutinho, 2015; Alho et al., 2019; Barbosa et al., 2020; Pandey et al., 2020; 38 Soares et al., 2021; Zhang et al., 2021). They also exhibit high cation exchange capacity (CEC) and moisture 39 retention among others properties (Hecht, 2003; Kämpf et al., 2003; Falcão et al., 2009). Charcoal content is 40 a key indicator of pre-Hispanic fire activity and sedentary occupation, which is evidence of the anthropic 41 origin of these soils (high confidence) (Hecht, 2017; Maezumi et al., 2018b; Alho et al., 2019; Barbosa et al., 42 2020; Iriarte et al., 2020; Montoya et al., 2020; Shepard et al., 2020). 43 44 Accumulation of organic residues and low intensity fires management are recognized as key elements for 45 ADEs formation. ADEs originating around settlements show a relatively high density of ceramic artefacts 46 and are named Terras pretas. They present a higher content of Ca and P than those originated from 47 agriculture activities which are known as Terras mulatas (Hecht, 2003). 48 49 There is a robust and growing body of research from different disciplines that gives high relevance to ADEs 50 in the region. It has been shown through archaeological and paleoclimatic data that Amazonian societies 51 which based their agricultural management on "Terras Pretas de Índio", were more resilient to the changing 52 climate due to increased soil fertility and water retention capacity (de Souza et al., 2019). Additionally, low 53 organic carbon degradability over long time periods, associated with high contents of charcoal or pyrogenic 54 carbon, makes these soils an important C sink (medium confidence: robust evidence, medium agreement) 55 (Lehmann et al., 2003; Guo, 2016; Trujillo et al., 2020), which is particularly relevant in an area like the 56 Amazon, that could change from a net carbon sinks to a net carbon source as a consequence of 57 anthropogenic climate change (Maezumi et al., 2018b). Do Not Cite, Quote or Distribute 12-95 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 2 The Indigenous agricultural practices which originated ADEs are thought to be associated with a more 3 sedentary agricultural model than the current slash and burn and shifting cultivation practices. Although this 4 is a controversial topic, as the precise definition of slash and burn and shifting cultivation is presently under 5 discussion (Hecht, 2003); several present-day local and Indigenous agricultural practices, including in-field 6 burning and nutrient additions from food processing and residue management, have been recognized as 7 promoting high organic carbon and nutrient soil contents similar to the ones found in ADEs (Hecht, 2003; 8 Winklerprins, 2009). 9 10 At present, ADEs are estimated to cover up to 3.2% of the Amazon basin and are highly valued for their 11 persistent fertility, becoming a key resource for sustainable agriculture for Amazon communities in a climate 12 change context (Altieri and Nicholls, 2013; Maezumi et al., 2018a; de Souza et al., 2019). Based on the 13 lessons learned from the Terras Pretas de Índio, some researches have proposed the development of 14 technologies to promote a new generation of anthropogenic soils (e.g., Kern et al. (2009); Lehmann (2009); 15 Schmidt et al. (2014); Bezerra et al. (2016); Kern et al. (2019)). Among the technologies based on ADEs 16 learnings Biochar, obtained by slow pyrolysis of agricultural residues, is the most explored application found 17 in literature (Mohan et al., 2018; Matoso et al., 2019; Amoah-Antwi et al., 2020). The dual purpose of 18 increased soil fertility and carbon sequestration is considered an important goal in order to develop 19 sustainable agriculture in a climate change context (Kern et al., 2019). 20 21 Preservation of the practices and knowledge associated with these soils is vital for sustainable agriculture in 22 a climate change scenario in the Amazon. It will greatly contribute to the preservation of valuable 23 Indigenous knowledge as well as the contribution to the development of new adaptation and mitigation 24 technologies among other unexplored solutions. 25 26 12.6.3 Towards a Metropolitan Water-related Climate Proof Governance (re)configuration? The case of 27 Lima, Perú 28 29 Lima-Callao Metropolitan City, capital of Perú is facing recurrent climate disasters showing lessons on 30 water-related climate-proof governance reconfiguration: 1) when disasters affect the poor and rich 31 population, dominant actors prioritize the integral city's resilience and development, and coordinate and 32 collaborate within a concertation manner across institutional levels and geographical scales (Hommes and 33 Boelens, 2017; Miranda Sara, 2021), even having different ideas, discourses, and power, recognizing that no 34 one single actor has enough power; 2) water-related climate change scenarios require comprehensive, 35 transverse, multi-sectoral, multi-scalar, multiple types of actor´s knowledge (expert, tacit, codified and 36 contextual embedded (Pfeffer, 2018) and transparent information to manage the tensions and even conflicts 37 when some knowledge is not shared or restricted particularly when lower risk perception and higher risk 38 tolerance are present; 3) a concertative (processes which involve a variety of actors and has become 39 mandatory in Peru) strategy to localize the climate action shows quicker, more effective and transparent 40 results (medium confidence, robust evidence, medium agreement) (Miranda Sara and Baud, 2014; Pepermans 41 and Maeseele, 2016; Siña et al., 2016; Miranda Sara et al., 2017). 42 43 Being the second driest city in the world, Lima is highly vulnerable to drought and heavy rainfall in the 44 nearby Andean highlands (Schütze et al., 2019). Located on the Pacific Coast with more than 10 million 45 inhabitants, suffers from both flooding, mudslides disasters and water stress, being more frequently affected 46 by heavy rain peak events (1970, 1987, 1998, 2012, 2014, 2015 and 2017) (very high confidence) (Mesclier 47 et al., 2015; Miranda Sara et al., 2016; French and Mechler, 2017; Vázquez-Rowe et al., 2017; Escalante 48 Estrada and Miranda, 2020). In addition to water unequal distribution in quantity and pricing, one million 49 inhabitants lack water connections (Ioris, 2016; Miranda Sara et al., 2017; Vázquez-Rowe et al., 2017) as a 50 result of a lack of long-term city planning and lack of integration with water and risk management. Climate 51 change scenarios were ignored or denied, particularly when the budget allocation for preventive actions was 52 necessary (high confidence) (Miranda Sara et al., 2016; Allen et al., 2017a). 53 54 In 2014, the Water Company (SEDAPAL) together with the Lima Metropolitan Municipality (LMM), ANA, 55 and other organizations agreed on a Lima Action Plan for Water (Schütze et al., 2019). The same year, the 56 Lima Metropolitan Municipality (LMM) approved the Climate Change Strategy defining adaptation and Do Not Cite, Quote or Distribute 12-96 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 mitigation measures (Miranda Sara and Baud, 2014), based on technical and scientific action research within 2 interactive, and iterative concertation multi-actor processes. 3 4 However, in 2015, municipal elections shifted Lima's and later Peru's political power to parties associated 5 with climate deniers at a high cost to the people, city infrastructure, and housing. Beginning of 2017, 6 buildings along rivers, ravines, and slopes suffered from floods, huaycos (mudslides), the whole city suffered 7 potable water cuts (Vázquez-Rowe et al., 2017) and vector-borne diseases affecting particularly the poorer 8 but also richer inhabitants. 9 10 "Coastal Niño", affected the whole country, as a consequence, in 2018, the Peruvian government passed the 11 Framework Law for Climate Change, Law No. 30754, a unique political decision, to assure the integration of 12 climate change concerns in public policies and investment projects. The law defines local governments 13 mandates on Local Climate Action Plans. The 2019 municipal elections brought new local authorities to 14 Lima and by 2020, 19 district municipalities developed their Adaptation Measures adopting the Metropolitan 15 Climate Change Strategy with support of Cities for Life Foro and GIZ (Foro Ciudades Para la Vida, 2021), 16 in 2021 LMM approved its Local Climate Change Plan (LCCP) and other 10 (out of 51 with Callao) 17 municipalities concluded the elaboration of their LCCP with support of the Global Covenant of Mayors and 18 the European Union. 19 20 The institutionalized culture of participation in Peru did lead to a broader concept of concertation, wherein 21 practices of collaborative planning were developed to allow actors to build up socially supported agreements, 22 decisions and take actions without losing sight of their principles. These processes have been applied to 23 reduce risks, to adapt and to anticipate uncertain and unknown futures; and introducing climate change 24 concerns within a complex political and institutional environment surrounded by corruption scandals 25 (Vergara, 2018; Durand, 2019) and growing political polarization. 26 27 Several processes have been set in motion to engage citizen participation and promote climate action 28 planning: 1) The LMM with the Climate Action Plan processes reopened the Climate Change Technical 29 Group of the Municipal Environmental Commission whose work ended in the approval of the Lima Local 30 Action Plan of Climate Change (MML, 2021), 2) The River Basin Council is developing the River Basin 31 Management Plan led by the National Authority of Water (ANA); 3) The Metropolitan Lima Urban 32 Development Plan is finalizing a citizen consultation, with the support of a high-level Consultation Group. 33 34 Such processes include strong discussions, conflicts, and the recognition of other's discourses and types of 35 knowledge, to build up scenarios that "visualize" and anticipate what might happen. These processes require 36 democratic, transparent, and decentralized institutions, providing clear mandates and strong political will to 37 support them, so the views of the poor and vulnerable are included, being able to make themselves heard, 38 even if their power remains limited (Chu et al., 2016). Opportunities for the reconfiguration of socio-political 39 and technological water governance are emerging based on socially supported agreements (Miranda Sara and 40 Baud, 2014; Miranda Sara, 2021). Although the water governance configuration faces the paradox that 41 current water demands of all users combined may no longer be feasible within ecological limits and future 42 climate change consequences (Miranda Sara et al., 2016; Schütze et al., 2019). 43 44 12.6.4 Strengthening Water Governance for Adaptation to Climate Change: Managing Scarcity and 45 Excess of Water in the Pacific Coastal area of Guatemala 46 47 Guatemala experiences high climate inter-annual variability now increased from the effect of climate change 48 (INSIVUMEH, 2018; Bardales et al., 2019). Impacts on human settlements, agriculture and ecosystems 49 result from both excess and reduced precipitation (high confidence) (Section 12.3.1.4). Guerra (2016) argues 50 that deficient integrated water resource management in the country is the main reason for those impacts. A 51 case in point is that of rivers Madre Vieja and Achiguate where an intense El Niño event triggered dryer 52 conditions and, in turn, a crisis and conflict that reached national proportions. Progress in local water 53 governance helped to solve that crisis and contributed to tackle challenges posed by reduced precipitation 54 and flood risk in southern Guatemala. 55 56 The ENSO event that started in November 2014 and ended in July 2016 (CIIFEN, 2016) has been the most 57 intense since records commenced in 1950 (NOAA, 2019). Its effects were felt in different parts of the world Do Not Cite, Quote or Distribute 12-97 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 and, Guatemala and the rest of Central America experienced an intense water scarcity due to a significant 2 reduction in rainfall (high confidence) (IICA, 2015; Scientific American, 2015). River flow in the dry 3 months is related to precipitation levels in the previous rainy season and thus, ENSO has an effect on river 4 flow rates. Two of the main rivers in the Pacific coast of Guatemala, Madre Vieja and Achiguate, dried out 5 completely at the beginning of 2016, triggering a nearly violent local conflict that caught attention at the 6 national level (Guerra, 2016; Gobernación de Escuintla et al., 2017). In addition to the severe drought, the 7 rivers dried because of over-extraction by multiple users (60 in the case of Madre Vieja). This had happened 8 before to a lesser extent in the last 20 years during the critical months of the dry season. Lack of regulation, 9 coordination mechanisms, information, and other elements of water governance was the root cause of the 10 problem, exacerbated by the drier conditions during the intense El Niño resulting in the intensification of an 11 existing conflict (high confidence) (Guerra, 2016). 12 13 Roundtables were set up to foster dialogue between numerous stakeholders including communities, agri- 14 export companies, governmental organisations, municipalities, all led by the local governor (Gobernación de 15 Escuintla et al., 2017). Agreements included: to keep a minimum of the rivers flowing all the way to the sea; 16 to set up a monitoring and verification system for levels of river flow; and to restore riparian forests. A 17 system was set up to monitor river flow in different points along the rivers on a daily basis in the dry season 18 using a simple WhatsApp-based system to communicate the warnings and monitor compliance. Four years 19 on, the rivers had not dried out and conflict was kept to a minimum. Rural communities can use rivers for 20 recreational purposes and for fishing all year round, whilst plantations (large and small) can use water for 21 irrigation (rationally) and keep producing. Similar schemes and interactions started happening in other rivers 22 in the Pacific coast of Guatemala, with positive results, particularly keeping the rivers flowing all through the 23 dry season as can be seen in the report of river flows for years 2017, 2018 and 2019 (ICC, 2019b). 24 25 A key actor in the improvement of water governance has been the Private Institute for Climate Change 26 Research (ICC). This is a unique initiative that was created in 2010 and is funded primarily by the private 27 sector of Guatemala to help the country advance in climate change mitigation and adaptation (Guerra, 2014). 28 The institute works alongside local governments, communities and private companies in several topics apart 29 from integrated water management. Its role is merely technical-scientific, being in charge of the water 30 monitoring system, generating data on weather and hydrology, and providing support to other stakeholders. 31 32 Local governance was also essential for the implementation of flood risk management actions (high 33 confidence). Guerra et al. (2017) explained how impacts were significantly reduced in the Coyolate river 34 watershed, also in the Pacific coast of Guatemala, thanks to flood protection that was designed and 35 implemented in a technical and integrated manner. This was a result of strong and active participation of 36 local communities, companies and the local municipality who demanded the central government to invest 37 effectively. The stakeholders provided some resources (financial and in-kind) and inspected the works. Some 38 flat areas of the lower Coyolate watershed used to flood annually causing economic damage for 39 communities. The areas covered by flood risk measures have not flooded which has avoided losses as well as 40 created conditions for investment to come and create jobs, improving life conditions for locals. Other 41 processes of participation and interaction between the authorities, the private sector and communities have 42 taken place in other watersheds for planning, action and investment for flood risk management. The ICC has 43 played a role by studying flood-prone areas, building capacities in communities, fostering public-private 44 coordination mechanisms, and providing much-needed technical assistance to local governments (ICC, 45 2019a). 46 47 Although some may argue that water governance is in the realm of development, it has made contributions in 48 reducing direct and indirect impacts of climate events and therefore, it can be seen as a key element for 49 climate adaptation (high confidence). 50 51 52 12.7 Knowledge Gaps 53 54 Data deficiencies and heterogeneity in quantity, quality and geographical bias in knowledge limit the 55 understanding of climate change, the evaluation of its impacts, and the implementation of adaptation and 56 mitigation measures (Harvey et al., 2018) in CSA. The number of publications is not representative with 57 respect to the sensitivity to climate change and vulnerability contexts of different subregions and sectors. Do Not Cite, Quote or Distribute 12-98 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 This lack of representation in the mainstream literature may lead to a bias and, therefore, an underestimation 2 of the overall climate-related impact for some CSA subregions (Sietsma et al., 2021). The reason for 3 relatively few quantitative studies might be the complexities of socio-demographic and economic factors, 4 and the lack of long-term and reliable data in these areas (Harvey et al., 2018), along with other social, 5 economic and technical constraints. 6 7 Most studies that assess vulnerability to climate change do not yet follow the concept adopted since the Fifth 8 Assessment Report (AR5) which separates exposure as an external variable (WGII AR5 Figure SPM 1) 9 (IPCC, 2014), and many still use the A and B system of climate change scenarios from AR4, as adoption of 10 the RCP models has been slow. There is still limited literature on severe risks and little specific and explicit 11 consideration of risk drivers in the region. Moreover, limits to adaptation and the effectiveness of adaptation 12 measures in CSA remain largely understudied. 13 14 The research of the interactions between climate change and socioeconomic processes is underdeveloped 15 (Barnes et al., 2013; Leichenko and O'Brien, 2019; Thomas et al., 2019). There is limited understanding of 16 the multilevel synergistic effects of climate change and other drivers including economic development from 17 household to country level (Wilbanks and Kates, 2010; Leichenko and Silva, 2014; Tanner et al., 2015a; 18 Carey et al., 2017). In the region, this deficit is deeper for sectors other than agriculture, water and food. 19 20 12.7.1 Knowledge Gaps in the Subregions 21 22 The knowledge gaps in the eight subregions are quite heterogeneous. In CA, climate change research is 23 notably insufficient in all sectors included in this report, considering that climatic change, variability, and 24 extremes are and will severely impact this subregion, and the vulnerability of the social and natural systems 25 is high. Data deficiencies must be overcome as renewed research on climate change updates models, 26 scenarios, and projected impacts across sectors and levels (i.e., household to country). In NWS, there is a 27 lack of studies on the relationships with increased fire events, and the impacts on the infrastructure of all 28 kinds, on certain lowland, marine and coastal ecosystems, and on ecosystem functioning and the provision of 29 environmental services. Experimental studies are rare, most necessary to identify critical ecological 30 thresholds to support the decision-making processes, linking glacier retreat to its consequences on 31 biodiversity and ecosystems, combined with different land-use trajectories. Complex interactions with 32 processes such as peace agreements in Colombia are yet to be studied (Salazar et al., 2018). In NSA, there is 33 still a limited amount of peer-reviewed literature, addressing the implications of climate change on 34 Indigenous cultures and their livelihoods. In SAM, further data are needed on the vulnerability of traditional 35 populations, impacts on water availability and soil degradation, risks to biodiversity and resilience of 36 ecosystems, attributed to climate change. 37 38 There is a knowledge gap about the likely impact of climate change on NES biodiversity, soil degradation, 39 and best adaptation measures. SES is the most urbanized sub-region of CSA, but there is a strong knowledge 40 deficits related to the design, implementation and evaluation of adaptation policy plans to climate change. 41 Forecasts related to risk prevention require new studies that address down-scaled climate change models 42 with concrete solutions to increase the city's resilience. In SWS, there is a lack of long-term studies 43 addressing climate change impacts in terrestrial, freshwater and marine ecosystems which is mainly due to 44 the lack of integrated observational systems. There is a lack of studies projecting future impacts of climate 45 change on the cryosphere, water resources, hazards, risks and disasters on natural and human systems. This 46 is mainly due to the lack of systematic documentation, analysis and evaluation of adaptation strategies 47 adopted, as well as their limitations and the lessons learned from maladaptation processes. There is low 48 evidence about transformational adaptation to climate change and systems resilience. In SSA, there is a need 49 for information related to vulnerability and impacts of the direct effects of future climate change on cities, 50 energy infrastructure and health. Also, there is a gap of knowledge about financing of climate change 51 adaptation in SSA. 52 53 12.7.2 Knowledge Gaps by Sector 54 55 12.7.2.1 Terrestrial and Freshwater Ecosystems and their Services 56 Do Not Cite, Quote or Distribute 12-99 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Advances on scientific knowledge on risks of climate change impact, vulnerability and resilience of 2 ecosystems is needed (Bustamante et al., 2020). Persistent climate change in tropical rainforest needs further 3 understanding, overall on the role of nutrients, deep-water availability and biodiversity. Further research is 4 needed to understand feedback to the climate systems of large-scale changes in the land surface in South 5 America biomes. The region has important freshwater Global-200 Ecoregions, including the Orinoco River 6 and Flooded Forests, Upper Amazon river and streams, and Amazon River and Flooded Forests being, 7 therefore, a priority for freshwater biodiversity conservation at a global scale (Manes et al., 2021) (Cross- 8 Chapter Paper 1; Figure 12.8). There is, however, a clear knowledge gap on the impacts of climate change on 9 freshwater biodiversity in the region (Cross-Chapter Paper 1.2.3; Manes et al., 2021) . Lastly, more 10 interdisciplinary research is needed regarding conservation strategies and stable financial resources focusing 11 on adaptation of ecosystems in the region (Mistry et al., 2016; Gebara and Agrawal, 2017; Ruggiero et al., 12 2019; To and Dressler, 2019). 13 14 12.7.2.2 Ocean and Coastal Ecosystems and their Service 15 16 There is an important lack of knowledge about the health state of the ocean and coastal ecosystems along 17 CSA (i.e., social-ecological data integration, poor sampling efforts, lack of information about the value of 18 ecosystem services, lack of information about ecosystems cover and distribution, lack of studies 19 about climate change perception and social concerns), including marine fisheries (i.e., landing statistics 20 not available, lack of reliable information on the scope of resource extraction, among others). Poor or absent 21 monitoring programs (physical, environmental and biological variables) that feed alert and surveillance 22 systems are missing for CSA. There is a general absence of a continuous line of scientific research or an 23 adequate baseline information about the impacts of climate change, as well as a continuous monitoring of the 24 adaptation plans adopted in ocean and coastal ecosystems which limit the formulation of adequate 25 conservation and management programs. When studies are performed, inadequate access to data limits the 26 analyses of the existing information making difficult to detect climate change trends and impacts, as well as 27 the development of effective adaptation strategies. 28 29 12.7.2.3 Water 30 31 As in other sectors and environmental systems, for the water sector there are important limitations in terms 32 of monitoring and data collection. High-quality, long-term hydrological data are unevenly available for 33 different subregions and limit a better understanding of changes in river runoff, lake or groundwater changes. 34 Groundwater data is particularly scarce. There are important gaps related to projections of water resources 35 for the future. Much of the current knowledge on future changes in water resources and water scarcity and 36 flood risks is based on information from global-scale studies because studies specific to this region are 37 scarce. Several elements which are important for integrated water resource management such as water 38 quality, water demand, privatization and other economic dynamics, and nutrient, pollutant and sediment flux, 39 are poorly known currently due to missing data and insufficient efforts to monitor them. 40 41 12.7.2.4 Food, Fibre and other Ecosystem Products 42 43 Integrative evaluation on impacts on food security, including agricultural production, distribution and access, 44 leading to adaptation strategies is limited within the region. Limited information regarding cost-benefit 45 analyses of adaptation in the food production sector is available in the region. It is also important to advance 46 in a better understanding of the adaptation effects to avoid maladaptation and promote site-specific and 47 dynamic adaptation options considering available technologies. Compiling and systematizing existing 48 scientific and local knowledge on the relationship between forest, land cover/use, and hydrological services, 49 is a gap to be filled, in a broader perspective in the region, that can contribute to provide recommendations 50 and inform restoration practices and policies. The literature also highlights widespread gaps between 51 farmers' information needs and services that are routinely available. There is evidence that when Climate 52 Information Services are constructed with farmer input and are targeted in a timely and inclusive manner, 53 they are a positive determinant of adaptation through the adoption of more resilient farm level practices. 54 However, currently assessments of the economic impact of Climate Information Services are scarce; hence 55 increased frequency of such studies is needed 56 Do Not Cite, Quote or Distribute 12-100 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 12.7.2.5 Cities, Settlements and Infrastructure 2 3 Despite the high level of urbanization in the region, studies on urban adaptation initiatives are still 4 underreported by municipalities and several practical results have not yet been demonstrated (Araos et al., 5 2016). It is particularly relevant to medium sized cities, as most of the literature and data available on 6 adaptation refers to the major capital cities. The potential of applying new resilient parameters in building 7 and land use regulation for adaptation is virtually underreported. The same can be said about the impact of 8 housing improvement and slum upgrading on climate resilience, even when initiatives are focused on 9 reducing environmental and climate risk. Also relevant in the region is a gap in research about NbS applied 10 to urban areas adaptation, as in the case of the urban forestry potential for adaptation (Barona et al., 2020). 11 Even though the importance of urban ecological infrastructure in providing ecosystem services, as flood 12 control, is reasonably documented, its practical application in urban planning in CSA is still limited 13 (Romero-Duque et al., 2020). Added to this is the lack of monitoring data on adaptation initiatives in 14 general, and in particular, on adaptation initiatives in water systems, that have already been implemented, 15 and its effects on risk reduction. Lack of monitoring data contributes to the lack of information about 16 maladaptation in urban areas and its consequences. Mobility and transport systems adaptation options are 17 virtually non-studied, while mitigation options receive a lot of attention. 18 19 12.7.2.6 Health and Wellbeing 20 21 There is a growing body of evidence that climate variability and climate change (CVC) cause harm to human 22 health in CSA. However, there is a lack of information about the current and future projected impact of CVC 23 events on overall illness and death in this region. It is challenging to attribute specific health outcomes to 24 CVC in models and field experiments due multiple factors including: 25 · lack of long-term high-quality health surveillance data 26 · multiple interacting infectious disease and chronic health issues 27 · mismatch in the spatial and temporal scales of CVC and health measurements 28 · complex climate and human system dynamics including nonlinear time-lags 29 · limited longitudinal data on non-climate factors that influence health outcomes (e.g., public health 30 interventions, migration of human populations, seasonal patterns in livelihoods). 31 32 The uncertainty inherent in predictive models also makes it challenging to expand current localized 33 knowledge on the impacts of infectious diseases associated with CVC to other regions or future climate 34 scenarios (UNEP, 2018). 35 36 Improved risk assessments based on better models and empirical research are needed to bridge the 37 knowledge gap and inform the design of adaptation strategies. A systematic multi-scalar analysis of the 38 impact of CVC on human health is needed across distinct social-ecological contexts. Data collection systems 39 need to be strengthened to accurately estimate the burden of mortality and morbidity from heat and extreme 40 events. The data deficit is a common problem in functioning civil registration and vital statistics systems, 41 including lack of information on causes of death (UNEP, 2018). In addition, there is a lack of consensus on 42 globally accepted and operational definitions for both climate-related extremes and exposures/outcomes. 43 44 For infectious disease (vector-borne and water-borne), the technology available to estimate current and 45 future risk areas is often limited by human or financial resource constraints in developing countries. There is 46 a geographical mismatch between the areas producing the technology and knowledge (in the global north), 47 and the areas most affected by CVC (in the global south). User-friendly tools that bring together climate and 48 health information--without the need for modelling or GIS expertise-- are needed for health sector decision 49 makers. 50 51 There is a lack of studies that assess the feasibility of health adaptation measures (see Section 12.5.10), thus 52 limiting the ability of decision makers to compare different health interventions and identify bottlenecks for 53 implementation. The growing field of implementation science could help to address barriers to 54 mainstreaming climate information in the health sector as an adaptation strategy. 55 56 Finally, there is an almost complete void of studies that address relationships of climate change with 57 wellbeing in CSA, broadly understood as including emotions and moods, satisfaction with life, sense of Do Not Cite, Quote or Distribute 12-101 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 meaning, and positive functioning, including the capacity for unimpaired cognitive functioning and 2 economic productivity (Section 7.1.4.1). 3 4 12.7.2.7 Poverty, Livelihood and Sustainable Development 5 6 Climate change is becoming a major obstacle in reducing poverty and overcoming poverty traps. There is a 7 need to better understand how poor and vulnerable communities are affected and the more effective ways to 8 prevent it. The large majority of the poor in the region are living in urban areas (UNDESA, 2019); urban 9 extreme poverty is increasingly more relevant, including the needs and priorities of informal settlements and 10 economies, but less studied within the interaction with climate change. There is little reporting of major 11 adaptation options implemented by or for vulnerable and poor urban dwellers (Ryan and Bustos, 2019; 12 Berrang-Ford et al., 2021). 13 14 Adaptation options are progressively being documented for poverty-related impacts in spite of the uncertain 15 context from climate impacts not being uniform across communities and the very local scale of the type of 16 adaptation responses needed (Miranda Sara et al., 2016; Rosenzweig et al., 2018; Dodman et al., 2019). 17 There is a huge gap in understanding how the poor are responding to climate change, what is needed to 18 support them, and the interconnections between development policies, poverty and risk reduction with 19 climate change actions (Ryan and Bustos, 2019; Satterthwaite et al., 2020). 20 21 The literature to assess the effectiveness of pro-poor or low-income adaptation options continues to be weak, 22 a very small proportion show results associated with adaptation efforts (Magrin et al., 2014; Berrang-Ford et 23 al., 2021). Without this kind of approach and in depth understanding there is the risk that top down climate 24 change adaptation options could reinforce poverty cycles and neglect cultural values, even eroding them 25 (Bartlett and Satterthwaite, 2016; Walshe and Argumedo, 2016; Allen et al., 2017a; Hallegatte et al., 2018; 26 Kalikoski et al., 2018; UN-Habitat, 2018). 27 28 The impacts of climate change on vulnerable groups are still understudied. There is little or no climate data 29 on remote mountain regions of CSA as well as research measuring the vulnerability of smallholders living 30 there, making it hard to assess the expected changes or the possible adaptation measures (Pons et al., 2017; 31 Donatti et al., 2019). 32 33 12.7.2.8 Cross Cutting Issues it the Human Dimension 34 35 There is a significant number of studies addressing the impacts of climate change on the Amazon forest 36 (Brienen et al., 2015; Doughty et al., 2015; Feldpausch et al., 2016; Rammig, 2020; Sullivan et al., 2020); 37 however, the assessment of tangible and intangible impacts of climate change on Indigenous Peoples 38 cultures and livelihoods in this forest, need to be further advanced (Brondízio et al., 2016; Hoegh-Guldberg 39 et al., 2018). 40 41 Studies on the perception of climate change in rural and urban populations throughout the region have 42 increased, but there is a lack of more specific research on the perception of specific groups, such as 43 economic or political actors, that influence public institutions and policies at the local, national level and 44 regional. 45 46 While studies on climate change gender differentiated impacts have grown over the past ten years in Central 47 and South America, studies on how gender intersects with other dimensions such as race, ethnicities, age or 48 rural/urban settings are still needed. This will help to further understand how gender inequalities are 49 connected to broader power structures of societies and, thus, to produce evidence on the importance of an 50 intersectional approach to climate change. 51 52 Regarding the relation of social movements and climate change adaptation, institutions and politics, two 53 major issues stem out: youth movements for climate change and the resistances, mainly urban, to climate 54 change adaptation policies. Little connection is found in research concentrating on resistance to climate 55 change adaptation policies and their interaction with the politics of place. Conflictivity related to climate 56 change is another under-studied issue. 57 Do Not Cite, Quote or Distribute 12-102 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Although there are several case studies on migrations and displacements caused by strong and immediate 2 climatic threats, such as hurricanes or floods, and on slow-onset impacts, such as droughts or temperature 3 increase, there are gaps in the attribution or relative weight of climate change in these processes. 4 5 Still important to note is that synergies between mitigation, adaptation, risk reduction and sustainable 6 development have not been jointly explored, which would better facilitate adaptation policy approaches. 7 8 There are critical knowledge gaps in the interlinkages between social and environmental dynamics that are 9 important for climate change adaptation, as in Andean forest landscapes. A salient knowledge gap in this 10 thematic area is the need to characterize how multilevel and multi-actor governance systems can enable 11 sustainable land management practices, including ecosystem restoration (Mathez-Stiefel et al., 2017). More 12 capacities are needed to increase the generation of relevant knowledge. Even small grant programs can 13 sustain research projects that target the linkages between knowledge and decision making at multiple scales 14 (Báez et al., 2020). 15 16 17 12.8 Conclusion 18 19 Central and South America (CSA) is a broadly heterogeneous region in its topography, ecosystems, urban 20 and rural territories, demography, economy, cultures and climates. The region relies on a strong agrarian 21 economy in which small producers and large industries participate, but also large industrialized urban 22 centres, oil production and mining. The region is one of the most urbanized of the world and home to many 23 Indigenous Peoples, some still in isolation, and exhibits one of the highest rates of inequality, which is a 24 structural and growing characteristic in CSA. Poverty and extreme poverty rates are higher among children, 25 young people, women, Indigenous Peoples, migrant and rural populations but urban extreme poverty is also 26 growing (very high confidence). Socioeconomic challenges are intensified by COVID crisis. Most countries 27 in CA are already ranked as the highest risk level worldwide due to its exposure combined to high 28 vulnerability and low adaptive capacity; the lack of climate data and proper downscaling are challenging the 29 adaptation process (high confidence). 30 31 Many extreme events are already impacting the region and projected to intensify including warming 32 temperatures and dryness, sea level rise, coastal erosion, ocean and lake acidification resulting in coral 33 bleaching, and increasing frequency and severity of droughts in some regions, with associated decrease in 34 water supply, that impact agricultural production, traditional fishing, food security and human health (high 35 confidence). In Central America (CA), 10.5 million people are living in the so-called Dry Corridor, a region 36 with an extended dry season and now more erratic rainfall patterns. A water crisis in Brazil affected the 37 major cities of the country between 2014 and 2016, becoming more frequent since then. Severe droughts 38 have also been reported in Paraguay and Argentina. In contrast, the urbanised areas of Northern South 39 America (NSA) are highly exposed to extreme floods (41% of urban population in the Amazon Delta and 40 Estuaries). Urban areas in the region are vulnerable for many reasons, notably high rates of poverty and 41 informality, poor and unevenly distributed infrastructure, housing deficits, and the recurrent occupation of 42 risk areas (high confidence). 43 44 Socio-ecological systems in the region are highly vulnerable to climate change, which acts in synergy with 45 other drivers such as land use change and deep socioeconomic inequalities. Most biodiversity-rich spots in 46 the region will be negatively impacted. The Cerrado and the Atlantic Forest (two important biodiversity-rich 47 spots where about 72% of Brazil's threatened species can be found) are exposed to different hazards 48 (extreme events, mean temperature increase) due to climate change. Many coastal areas and its concentrated 49 urban population and assets are exposed to sea level rise. Climate change is threatening several systems 50 (glaciers in the Andes, coral reefs in Central America, the Amazon forest) that are already approaching 51 critical conditions under risk of irreversible damage. 52 53 Extreme heat, droughts and floods will seriously affect CSA terrestrial and freshwater ecosystems. The high 54 poverty level increases the vulnerability to droughts, both in cities and rural areas, where people already 55 suffer from natural water scarcity (high confidence). The conversion of natural ecosystems to other land uses 56 exacerbate the adaptation challenges. Indigenous knowledge and local knowledge play an important role in 57 adaptation but are also threatened by climate change (high confidence). Ecosystem-based Adaptation (EbA) Do Not Cite, Quote or Distribute 12-103 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 and Community-based Adaptation (CbA) have increased since AR5, with emphasis on freshwater 2 ecosystems and forests, including protected areas. Inadequate access to finance and technology are widely 3 identified as adaptation barriers (high confidence). 4 5 Many impacts in the economy are expected from climate change. Subsistence farmers and urban poor are 6 expected to be the most impacted by droughts and variable rainfall in the region (high confidence). The 7 increasing water scarcity is and will continue to impact food security, human health and well-being. The 8 impacts of the many landslides and floods affect mainly the urban poor neighbourhoods and are responsible 9 for the majority of the deaths related to natural disasters. Sea-level rise and intense storm surges are expected 10 to impact the tourism and industry in general. Internal and international migrations and displacements are 11 expected to increase (high confidence). Climatic drivers such as droughts, tropical storms and hurricanes, 12 heavy rains and floods, interact with social, political, geopolitical and economical drivers (high confidence). 13 14 The common patterns and problems, however, highlight also the possibilities for collaboration and learning 15 among the countries and institutions in the region in order to strengthen the interface between knowledge and 16 policy in climate change adaptation. All countries in the region have submitted their first and updated NDC, 17 and many have published their NAP, establishing priorities and formulating their own policies to cope with 18 climate change. 19 20 Various adaptation initiatives have been initiated in different sectors, focused on reducing poverty, 21 improving livelihood and achieving sustainable and resilient development. There is an increase in planned 22 and autonomous initiatives, led by community, government or the combination of both, engineering or 23 Nature-based Solutions (NbS). Climate smart agriculture is an effective option, in several conditions and 24 regions, to mitigate negative impacts of climate change. Disaster reduction solutions are increasingly used, 25 such as Early Warning Systems (EWS). Many and diverse initiatives are still poorly reported and evaluated 26 in the scientific literature, leading to challenges in its assessment and improvements, including the 27 consideration of the tacit Indigenous knowledge and Local Knowledge (IK and LK). The lack of climate data 28 and proper downscaling, weak governance, hindrance on financing, and inequality are constraining the 29 adaptation process (high confidence). 30 31 Adaptation measures have been increased and improved since AR5 in ocean and coastal ecosystems. The 32 majority of these measures are focused on EbA application through the application of protection and 33 recovery of already impacted ecosystems. Another battery of measures is focused on the management and 34 sustainability of marine resources subjected to fisheries, however these measures are not assessing current 35 and future climate change impacts but they are focused on decreasing the impact of other non-climate factors 36 such as overfishing or pollution. To date, along CSA there is an important lack of long-term research 37 addressing ocean and coastal ecosystems health and their species through continuous monitoring which is 38 one of the main barriers to adaptation. The number and type of adaptation measures for ocean and coastal 39 ecosystems and their contributions to humans are highly different among CSA countries which highlight in 40 number those measures related to increase the scientific research and monitoring followed by the 41 conservation of biodiversity, and changes in legislation (high confidence). On the other hand, those measures 42 that include the changes in financing (an important barrier) or the incorporation of traditional knowledge are 43 not always considered in national adaptation plans by CSA countries. 44 45 In the water sector a lack of systematic analysis and evaluation of adaptation measures prevail, although 46 important progress has been made since the AR5 in terms of understanding interlinkages between climate 47 change, human vulnerabilities, governance, policies and adaptation success (high confidence). NbS, Payment 48 for Ecosystems Services (PES), integrated water resource management, and integration of IK and LK have 49 proven potential of success, in particular if adopting approaches with inclusive negotiation formats for water 50 management with clear, just and transparent rights and responsibilities. 51 52 Climate change poses several challenges to the agri-food sector, impacting the agricultural production and 53 productivity, and posing at risk the food and nutritional security and the economy (high confidence). 54 Adapting agriculture while conserving the environment is a challenge for a sustainable and resilient food 55 production (high confidence). Adaptation in the region presents persistent barriers and limitations (Table 56 12.8), associated with investments and knowledge gaps (medium confidence). Climate change urges to Do Not Cite, Quote or Distribute 12-104 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 advance in initiatives to improve education, technology and innovation of farming systems in the CSA 2 region. 3 4 Urban adaptation is limited by financing constraints, weak intersectoral and multilevel governance, and 5 deficits in the housing and infrastructure sectors, the overcoming of which is an opportunity for 6 transformative adaptation (high confidence). Short-term interventions are prevailing over long-term planning 7 (high confidence). Adaptation experiences in planning, land use and building regulation, urban control 8 systems and risk management have taken place throughout the region. Initiatives in social housing are 9 reducing risk, overcoming urgent deficits, but also adding to a transformative adaptation pathway (high 10 confidence). Hybrid (green-grey) infrastructure has been adopted for better efficiency in flood control, 11 sanitation, water scarcity and landslide prevention and coastal protection (high confidence). NbS including 12 green infrastructure and EbA are increasing in urban areas (high confidence), although isolated engineering 13 solutions are still widely practiced. The integration of transport and land use plans and the improvement of 14 public transport are key to urban adaptation; mitigation prevails over adaptation in the sector (high 15 confidence). 16 17 There is a growing body of evidence that climate variability and climate change are causing harm to human 18 health in CSA ­ including the increasing transmission of vector borne and zoonotic diseases, heat stress, 19 respiratory illness associated with fires, food and water insecurity associated with drought, among others 20 (medium confidence). In response, countries in the region are developing innovative adaptation strategies to 21 inform health decision making such as integrated climate-health surveillance systems and observatories, 22 forecasting of climate-related disasters, and epidemic forecast tools. However, institutional barriers (limited 23 resources, administrative feasibility, and political mandates) need to be addressed to ensure the sustained 24 implementation of adaptation strategies (high confidence). 25 26 Poor and vulnerable groups evidence limited political influence, fewer capacities and opportunities to 27 participate in decision and policy making being less able to leverage government support to invest on 28 adaptation measures (very high confidence). Participatory processes are developing adaptation measures 29 strengthening local capacities; literature assessing the success of such initiatives remains limited. Limits to 30 adaptation include access to land, territory and resources, labour and livelihood opportunities, knowledge 31 gaps and poor multi actor coordination. Social organization, participation and governance reconfiguration are 32 essential for building climate resilience (very high confidence). 33 34 Social organization, participation, governance, education and communications to increase perception and 35 knowledge, are essential for building the resilience to adapt and overcome expected and unexpected climate 36 impacts (very high confidence). The focus on inclusion and enrolling of the full range of actors in adaptation 37 processes, including vulnerable populations, has shown good results in the region (high confidence). 38 However, existing poverty and inequality, unbalances on power relations, corruption, weak governance and 39 institutions, structural problems and high levels of risk tolerance may reinforce poverty and inequality cycles 40 (high confidence). In addition, the continued exposure of critical infrastructure and valuable assets are signs 41 of persisting maladaptation. 42 43 The development model prevailing in the region for the last decades has proven to be unsustainable, with the 44 emphasis on financial sources based on natural resource depletion and extraction and the persistence and 45 growing inequality. It is well recognized that climate adaptation measures, if carefully selected considering 46 the coupled human-environment systems, will provide significant contributions to the sustainable 47 development pathways of the region and to achieve the sustainable development goals (SDG) if implemented 48 together with comprehensive strategies to reduce poverty, inequality, and risks (high confidence). Adaptation 49 and the construction of resilience offer not only an opportunity to reduce climate change impacts, but also 50 the opportunity to reduce inequality and development gaps, to achieve dynamic economies, and to regulate 51 the sustainable use and transformation of the territory. 52 53 54 [START FAQ 12.1 HERE] 55 Do Not Cite, Quote or Distribute 12-105 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 FAQ 12.1: How are inequality and poverty limiting options to adapt to climate change in Central and 2 South America? 3 4 Poverty and inequality decrease human capacity to adapt to climate change. Limited access to resources 5 may reduce the ability of individuals, households and societies to adapt to the impacts of climate change and 6 variability because of the narrow response portfolio. Inequality limits responses available to vulnerable 7 segments as most adaptation options are resource-dependent. 8 9 Though poverty in Central and South America has decreased over the last 12 years, inequality remains as a 10 historic and structural characteristic of the region. In 2018, 29.5% of Latin America's population (including 11 Mexico) were poor (182 million) and 10.2% were extremely poor (63 million), more than half of them living 12 in urban areas. In 2020, due to COVID crisis Gini coefficient projection of increases is ranging from 1.1% to 13 7.8%, poverty increased to 33.7% (209 millions) and extreme poverty to 12.5% (78 millions). 14 15 Poor populations have little or no access to good quality education, information, health systems and financial 16 services. They have lower chances to access resources such as land and water, good quality housing, risk 17 reducing infrastructure and services such as running water, sanitation and drainage. Their lack of political 18 clout and endowments limit their access to assets for withstanding and recovering from shocks and stresses. 19 Poverty, inequality, and high vulnerability to climate change are inter­related processes. Poor populations 20 are highly vulnerable to impacts from climate change and are usually located in areas of high exposure to 21 extreme events. The constant loss of assets and livelihoods both in urban and rural areas drives communities 22 into chronic poverty traps, exacerbating local poverty cycles and creating new ones. 23 24 For instance, climate­related reduced yields in crops, fisheries, and aquaculture have a substantial impact on 25 the livelihoods and food security of families and affect their options to cope and adapt to climate change and 26 variability. The impact of climate change in agriculture for Central and South America depends on 27 determinants such as availability of natural resources, access to markets, diversity of inputs and production 28 methods, quality and coverage of infrastructure, as well as socioeconomic characteristics of the population. 29 Impacts from climate change on small­scale farmers compromise the livelihoods and food security of rural 30 areas and consequently the food supply for urban areas. 31 32 Governments in the region have implemented several poverty­reduction programs. However, policies of 33 income redistribution and poverty alleviation do not necessarily improve climate risk management, hence 34 complementary policies integrating both social and material conditions are required. A study in Northern 35 Brazil shows risk management strategies for droughts and food insecurity did not change poverty incidences 36 between 1997­1998 and 2011­2012. Major shocks, such as climate and weather extreme events (e.g., floods, 37 heavy rains, droughts, frosts), reduce and destroy public and private property. For instance, the ENSO event 38 of 2017 in Peru caused losses estimated between USD 6 to 9 billion, affected more than a million inhabitants 39 and generated 370,000 new poor. In total, losses by unemployment, deaths, destruction and damage of 40 infrastructure and houses were around 1.3% of the Gross Domestic Product of Peru. 41 42 Low public expenditure on social infrastructure (health, education etc.), ethnic discrimination and social 43 exclusion reduce healthcare access, leaving poor people in entire regions mostly undiagnosed or untreated. In 44 a context of privatization policies of health care systems, research shows marginal people lack identifying 45 documents needed to access public services in Buenos Aires (Argentina), Mexico City (Mexico) and 46 Santiago de Chile (Chile), some of the most developed cities in the region. Consequences of this situation are 47 under reporting, low diagnosis, and low treatment of diseases such as vector­borne diseases such as dengue 48 and risk of diarrheal diseases originated by frequent floods in Amazonian riverine communities. Bias on 49 reporting access to health­care and incidence of diseases in marginal populations are usually region­ 50 dependent. For example, in Brazil's Amazonian North in 2018, there were 2.2 medical doctors per 1000 51 inhabitants, while 4.95 medical doctors per 1000 inhabitants in São Paulo and 9.52 doctors in Santa Catarina. 52 Another example: pregnant women in remote Amazonian municipalities receive less prenatal care than 53 women in urban areas. These social inequities underlie systemic biases in health data­quality hindering 54 reliable estimation of disease burdens such as distribution of disease or birth and death registrations. For 55 Example, in Guatemala alternative Indigenous healthcare systems are responding to local needs by Mayan 56 communities. However, this remains unrecognized. The existence of health institutions based on Indigenous Do Not Cite, Quote or Distribute 12-106 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 knowledge can reinforce the lack of universal coverage by central government healthcare, addressing the 2 miscalculation of morbidity, mortality, and cause­of­death among disadvantaged groups. 3 4 Inequality, informality and precariousness are particularly relevant barriers for adaptation. A significant part 5 of the construction sector in the region is informal and does not follow regulations for land use and 6 construction safety codes, and there is a lack of public strategies for housing access. Adaptive construction is 7 based upon up­to­date regulation and codes, appropriate design and materials, and access to infrastructure 8 and services. Decreasing inequality and eradicating poverty are crucial for achieving proper adaptation to 9 climate change in the region. Some experiences to fight poverty such as savings groups, microfinance for 10 improving housing or assets and community enterprises may also support specific adaptive measures. These 11 mechanisms should be widely accessible to poor groups and be complemented by comprehensive poverty 12 alleviation programs that include climate change adaptation. 13 14 [END FAQ 12.1 HERE] 15 16 17 [START FAQ 12.2 HERE] 18 19 FAQ 12.2: How have urban areas in Central and South America adapted to climate change so far, 20 which further actions should be considered within the next decades and what are the limits of 21 adaptation and sustainability? 22 23 Cities are becoming focal points for climate change impacts. The rapid urbanization in Central and South 24 America, together with accelerating demand for housing, resource supplies and social and health services, 25 put pressure on the already stretched physical and social infrastructure. In addition, migration is negatively 26 affecting the opportunities of cities to adapt to climate change. 27 28 Central and South America is the second most urbanized region in the world after North America with 81% 29 percent of its population being urban. 129 secondary cities with 500,000 inhabitants concentrate half of the 30 region's urban population (222 million). Another 65 million people live in megacities over 10 million each. 31 The population migrates among cities, resulting in more secondary cities and creating mega regions and 32 urban corridors. 33 34 Rapid growth in cities has increased the urban informal housing sector (e.g., slums, marginal human 35 settlements and others), which increased from 6 to 26 percent of the total residences from 1990 to 2015. 36 Coastal areas in Central and South America increasingly concentrate more urban centres. Researchers 37 indicate that between 3 to 4 million inhabitants will experience coastal flooding and erosion from sea­level 38 rise in all emission scenarios by 2100 considering Southern America alone. 39 40 A study on cities with more than 100,000 inhabitants shows the number of coastal cities significantly 41 increased from 42 to 420 between 1945 and 2014; they are located close to fragile ecosystems such as bays, 42 estuaries and mangrove forests, resulting in higher concentrations of population and economic activities. 43 This process degraded the ability of coastal ecosystems, such as mangroves, to reduce risks and provide 44 essential ecosystem services which help to prevent coastal erosion or maintain fish stocks. Moreover, it 45 reduced ports, tourism, along with income opportunities. 46 47 Climate change impacts on cities in Central and South America are strongly influenced by El Niño Southern 48 Oscillation (ENSO) associated with an increase of more extreme rainfall events. Urban areas are increasingly 49 dealing with floods, landslides, storms, tropical cyclones, water stress, fires, spread of vector­borne and 50 infectious diseases, damaging infrastructure, economic activities, built and natural environments and the 51 population's overall well­being. 52 53 Glacier retreat in the mountains will affect water runoff and water provision to Metropolitan cities such as 54 Lima, La Paz, Quito and Santiago who rely on rivers that originate in the high Andes. Lima, the second 55 driest capital city in the world, is vulnerable to drought and heavy rain peak events associated with climate 56 change. In Bogota lower precipitations and a tendency of increasing extreme events are expected in the Do Not Cite, Quote or Distribute 12-107 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 coming decades. Hence, the protection of fragile ecosystems such as `paramo' (fields at 3000 to 4000 2 m.a.s.l.) will be crucial for water supply to the city. 3 4 Sea level rise impacts cities located in low elevation coastal zones, not only because of direct coastal 5 flooding, coastal erosion and subsidence; but also because it aggravates the impact of storm surges, heat 6 wave energy and saltwater intrusion. 68 percent of the population of Surinam and 31 percent of the 7 population in Guyana live below 5 meters above sea level, while many sectors of Georgetown, the capital of 8 Guyana, are below sea level. Floods with increased frequency and severity of storm surges will also impact 9 the Rio de la Plata estuary and lower delta of the Parana River where Metropolitan Buenos Aires is located. 10 11 Over 80 percent of losses associated with climate-related risks concentrate in urban areas, and between 40 12 and 70 percent losses occur in cities with less than 100,000 inhabitants, most probably as a result of limited 13 capacities to manage disaster risks and low level of investments. 14 15 Despite consistent political and economic barriers, many cities in the region have adopted sustainable local 16 development agendas, which work to address a balanced urban development. The shortcomings of poor 17 development patterns are still very present in the cities and present important obstacles to adaptation 18 investment, as public investment in basic needs (mainly housing and sanitation) must be prioritized. 19 20 Cities struggle to address the immediate needs of their population while addressing longer­term needs 21 associated with climate adaptation, emissions reduction and sustainable development. Some cities are 22 moving forward to transformative adaptation, addressing drivers of vulnerability, building robust systems 23 and anticipating impacts. Besides government­led adaptation planning and action, individuals, communities 24 and enterprises have been incrementally adapting to climate changes autonomously over time. Municipalities 25 from Argentina, Peru, Chile, Equator, Brazil and Costa Rica are developing and implementing their Local 26 Climate Action Plans, experimenting and displaying best practices in adaptation. Both anticipatory 27 adaptation measures­choosing safe locations, building structurally­safe houses, choosing elevated places to 28 store valuables, building on stilts­and reactive adaptation measures are used; the latter incorporating 29 measures such as relocation, stabilization of slopes, afforestation, and greening of riverbanks. With 30 variations, these cities have included mechanisms to work across sectors and actors understanding it is 31 collective planning and actions, which will ensure that long term programs continue independently of 32 particular city administrations. 33 34 Cities are interconnected systems operating beyond administrative boundaries. Improved collaboration and 35 coordination is needed for integrated responses. Aside from good planning, cities need access to external 36 adaptation funds. Climate change adaptation requires long­term funding and investments, which are beyond 37 cyclical political terms. It is key to re­think how to make international adaptation funds reach cities and 38 innovate. For example, member cities of Global Covenant of Mayors in the region, together with Cities for 39 Life Forum in Peru, the Red Argentina de Municipios por el Cambio Climático (RAMCC), the Capital Cities 40 of the Americas facing Climate Change (CC35) and others, are pursuing this goal and applying directly for 41 international grants. New funding sources are required to help local governments and civil society. Cities and 42 locally driven adaptation initiatives can be funded by national governments and international organizations. 43 44 [END FAQ 12.2 HERE] 45 46 47 [START FAQ 12.3 HERE] 48 49 FAQ 12.3: How do climatic events and conditions affect migration and displacement in Central and 50 South America, will this change due to climate change, and how can communities adapt? 51 52 Migration and displacements associated with climatic hazards are becoming more frequent in Central and 53 South America, and it is expected they will continue to increase. These complex processes require 54 comprehensive actions in the places of origin and reception, both to improve adaptation in the more affected 55 places, and the conditions of the mobilizations. 56 Do Not Cite, Quote or Distribute 12-108 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Migration of individuals, families and groups, voluntary and involuntary, is common in Central and South 2 America. People migrate nationally and internationally, temporarily or permanently, predominantly from 3 rural areas ­ often immersed in poverty ­ to urban areas. Common social drivers of migration in the region 4 are the economy, politics, land tenure and land management change, lack of access to markets, lack of 5 infrastructures, and violence; environmental drivers include loss of water, crops and livestock, land 6 degradation and sudden or gradual onset of climate hazards. 7 8 The increasing frequency and magnitude of droughts, tropical storms, hurricanes, and heavy rains producing 9 landslides and floods, have amplified internal movements, overall rural to urban. For instance, rural to urban 10 migration in Northern Brazil, or international migration from Guatemala, Honduras and El Salvador to North 11 America, are partly a consequence of prolonged droughts, which have increased the stress of food 12 availability in these highly impoverished regions. Diminished access to water is also a result of privatization 13 of that resource. In Central America, the majority of migrants are young men, reducing the labour force in 14 the places of origin. However, the migrants send back substantial amounts of money that have become the 15 main source of foreign exchange for their countries, and the main source of income for their families. 16 17 As poor people have less resources to adapt to changing conditions, they are usually the most impacted by 18 climate hazards, as they are already struggling to survive under normal conditions. These populations are the 19 most susceptible to migration, chiefly because of the loss of their livelihoods, their precarious housing and 20 settlements and the lack of money and international aid. Other important factors are the minimal 21 governmental support and assistance through social safety nets and extension services, the scarcity and low 22 quality of education and health services, the isolation and marginality, and the insecurity of land rights. 23 These same conditions, though, may hinder their mobility or even render them immobile. Nevertheless, in 24 some cases, despite worsening conditions, people decide not to move. 25 26 The magnitude and frequency of droughts and hurricanes are projected to keep increasing by 2050, which 27 may force millions of people to leave their homes. Climate models show some dry regions will become even 28 dryer in the coming decades, increasing the stress on small farmers who rely on rainfall to water their fields. 29 Glacier retreat and water scarcity are becoming strong drivers of migration in the Andes. Sea level rise 30 influences activities such as fishing and tourism, which will foster further migration. In Brazil, at least 0.9 31 million more people will migrate inter­regionally under future climate conditions. 32 33 Addressing migration and displacement requires diverse interventions: in dry regions it is recommended to 34 improve the water management in the places of origin of migration, including storage, distribution and 35 irrigation. Wet regions, lowlands, and floodplains will benefit from preventing construction on areas prone to 36 landslides and flooding. Government and international aid are also important for improving people's options 37 to adapt and enhance their resilience to climate impacts. In northern Brazil, for example, government 38 financial support has significantly reduced the migration caused by droughts. Between Guatemala and 39 Canada there is a temporary migration program to bring in migrant workforce during the harvest season. The 40 United States is also increasing these types of legal temporary migration. 41 42 [END FAQ 12.3 HERE] 43 44 45 [START FAQ 12.4 HERE] 46 47 FAQ 12.4: How is climate change impacting and expected to impact food production in Central and 48 South America in the next 30 years and what effective adaptation strategies are and can be 49 adopted in the region? 50 51 Agriculture is a fundamental sector to the development of societies from the economic and social 52 perspectives, and so it is a major component of the adaptive strategies for Central and South America 53 countries. Implementation of sustainable agriculture practices such as improved management on native 54 grasslands or agroforestry systems for crop and livestock production, can increase productivity while 55 improving adaptability. 56 Do Not Cite, Quote or Distribute 12-109 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Over the last two decades, countries throughout Central and South America have been developing rapidly. 2 The agriculture sector is fundamental to this development from economic and social perspectives. Some 3 countries of the region are amongst major food exporters in the world: 4 · Corn: three of the top ten exporters are Brazil, Argentina and Paraguay; 5 · Soybean exports: Brazil and Argentina are among the top five; Paraguay and Bolivia figure within 6 the twelfth; 7 · Coffee exports: five of the top ten export countries are Brazil, Colombia, Honduras, Peru and 8 Guatemala; 9 · Fruits: two of the top 10 fresh fruit export countries are Chile and Ecuador, 10 · Fishmeal exports globally are led by Peru, Chile and Ecuador; 11 · Beef: four of the top exporting countries are from this region ­ Brazil, Argentina, Uruguay and 12 Paraguay. 13 14 CSA is one of the regions with the highest potential to increase food supplies particularly to more densely 15 populated regions in Asia, Middle East and Europe. Better understanding the impact of the economy on the 16 environment and the contribution of the environment to the economy, is critical to identify opportunities for 17 innovation and promoting activities that could lead to sustainable economic growth without depleting natural 18 resources and increasing sensitivity to climate change and climate variability. The consideration of food as 19 a commodity instead of a common resource, leads to the accumulation of under-priced food 20 resources at the expense of natural capital. Without serious emissions reduction measures, climate 21 models project an average 1 to 4°C increase in maximum temperatures, and a 30 percent decrease in rainfall 22 towards 2050, across Central and South America. Tropical South America is projected to warm at higher 23 rates than the southern part of South America. Given these circumstances, some regions in Central and South 24 America (Andes region and Central America) will just meet, or fall below, the critical food supply/demand 25 ratio for their population. Meanwhile, the more temperate part of South America in the south is projected to 26 have agricultural production surplus. The challenge for this region will be to retain the ability to feed and 27 adequately nourish its internal population as well as making an important contribution to the food supplies 28 available to the rest of the world. 29 30 The Nationally Determined Contributions (NDCs) of most of the countries of Central and South America 31 expressly included agriculture as a major component of their adaptive strategy. From the recommendations 32 presented, five general adaptive themes, or imperatives, emerge: 1) inclusion of climate change projections 33 as a key element for Ministries of Agriculture and research institutes in their decision-making processes; 2) 34 support research and adoption of drought- and heat-tolerant crop varieties; 3) promotion of sustainable 35 irrigation as an effective adaptive strategy; 4) recovery of degraded lands and sustainable intensification of 36 agriculture to prevent further deforestation; and 5) implementation of climate smart practices and 37 technologies to increase productivity while improving adaptability. 38 39 Climate smart-practices provide a framework to operationalize actions aimed at understanding synergies 40 among productivity, adaptation and mitigation. Significant amount of evidence supports the potential for 41 climate smart-practices technologies to produce such triple wins as natural pastoral systems in the southern 42 region of South America. Such systems allow the combination of food production and environmental 43 sustainability. The production of meat based on native grasslands with grazing management that optimizes 44 forage allowance can achieve high production levels, while providing multiple ecosystem benefits. Optimal 45 forage allowance means offering the animals enough forage in order to meet requirements and while 46 avoiding overgrazing. This management practice simultaneously increases productivity, reduces greenhouse 47 gas emissions while improving soil carbon sequestration, and minimizes other environmental impacts such 48 as excess of nutrients, fossil energy use, and biodiversity loss. Pastoral farming systems that manage grazing 49 and feeding efficiently, are an example of integration between food security, environmental conservation and 50 nature-based adaptation to climate change. 51 52 Agroforestry systems are present in the tropical region of Central and Southern America. Trees are present in 53 a large part of the agricultural landscape of this region, either dispersed or in lines; supporting the production 54 of coffee, cocoa, fruits, pastures and livestock in various agroforestry configurations. In Central America, 55 shade-grown coffee reduces weed control and improves quality and taste of the product. Agroforestry uses 56 nitrogen-fixing trees (Leguminosae), such as Leucaena in Colombia, Inga in Brazil, to restore soil nitrogen 57 fertility. Tropical forest soils are generally nutrient-poor and unsuited to long-term agricultural use. Land Do Not Cite, Quote or Distribute 12-110 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 converted to agriculture by cutting and burning natural vegetation tends to remain productive for only a few 2 years. Agroforestry and the so called silvopastoral systems, which incorporate trees into crop and livestock 3 systems, have been shown to make a dramatic impact on the maintenance and restoration of long term 4 productivity in agricultural landscapes, including degraded and abandoned land. Agroforestry systems can 5 provide major benefits through enhanced food security, stronger local economies, and increased ecosystem 6 services such as carbon storage, regulation of climate and water cycles, control of pests and diseases, and 7 maintenance of soil fertility. Because of these multiple goods and services, agroforestry practices are 8 considered one of the key strategies for the development of climate smart agriculture. 9 10 [END FAQ 12.4 HERE] 11 12 13 [START FAQ 12.5 HERE] 14 15 FAQ 12.5: How can Indigenous knowledge and practices contribute to adaptation initiatives in 16 Central and South America? 17 18 Indigenous Peoples have knowledge systems and practices that allow them to adapt to many climatic 19 changes. Adaptation initiatives based on Indigenous knowledge and practices are more sustainable and 20 legitimate among local communities. It is important to build effective and respectful partnerships among 21 Indigenous and non­indigenous researchers to co­produce climate relevant knowledge to enhance 22 adaptation planning and action in the region. 23 24 There are 28 million Indigenous Peoples in Central and South America (around 6.6% of the whole 25 population of the region). They belong to more than 800 groups living in territories covering a wide range of 26 ecosystems ­ from drylands to tropical forests to savannahs, coasts to mountains ­ and that share the land 27 with many other cultural and ethnic groups. In the region, Indigenous Peoples are often categorized as a 28 group highly vulnerable to climate change as they are frequently affected by socioeconomic inequalities and 29 the dominance by external powers. They often experience internal and external pressures over their 30 communal lands in forms of pollution, oil and mining, industrial agriculture, and urbanization. On the other 31 hand, it is important to recognize that Indigenous Peoples have knowledge systems and practices that allow 32 them to adapt to many climatic changes. Increasing scientific evidence shows that adaptation initiatives 33 based on Indigenous knowledge and practices are more sustainable and legitimate among local communities. 34 35 The wide range of adaptation practices based on Indigenous knowledge in the region include, among others: 36 increasing species and genetic diversity in agricultural systems through community seed exchanges; 37 promotion of highly diverse crop systems; ancient systems to collect and conserve water; fire prevention 38 strategies; observing and monitoring changes in communal ecological­agricultural calendar cycles; 39 recognizing changes in ecological indicators like migration patterns in birds, behaviour of insects and other 40 invertebrates and phenology of fruit and flowering species; and systematization and knowledge exchange 41 among communities. These practices represent a valuable cultural and biological heritage. 42 43 The Kichwa in the Ecuadorian Amazon cultivate Chakras (plots) within the rainforest. These plots combine 44 crops and medicinal herbs for both self­consumption and selling. Similar systems, like the Chakras in the 45 high Andes, the Milpas in Central America, and the Conucos in northern South America have been resilient 46 to social and environmental disturbances due to their outstanding agrobiodiversity (more than 40 species and 47 varieties can be present in one plot), microhabitat management and the associated knowledge and 48 institutions. 49 50 Traditional fire management among Indigenous Peoples of Venezuela, Brazil and Guyana is another 51 adaptation strategy based on a fine­tuned understanding of environmental indicators, associated with their 52 culture and worldviews. In these countries, Indigenous lands have the lowest incidence of wildfires, 53 significantly contributing to maintaining and enhancing biodiversity. These traditional practices have helped 54 to prevent large­scale and destructive wildfires, reducing the risk from rising temperature and dryness due to 55 climate change. 56 Do Not Cite, Quote or Distribute 12-111 Total pages: 181 FINAL DRAFT Chapter 12 IPCC WGII Sixth Assessment Report 1 Traditional agriculture of Mapuche Indigenous Peoples in Chile includes a series of practices that result in a 2 system more resilient to climate and non­climate stressors. Practices include water management, native seed 3 conservation and exchange with other producers (trafkintu), crop rotation, polyculture, and tree­crop 4 association. Similar practices can be found in Mayan communities in Guatemala at the other end of the 5 subcontinent. 6 7 Despite the increasing recognition and integration of Indigenous knowledge in adaptation practices and 8 policies in the region, important barriers for a more effective and transformative integration remain. Some of 9 the most relevant barriers include limited participation of Indigenous Peoples and local communities in 10 adaptation planning and the lack of sufficient consideration of non­climatic socioeconomic drivers of 11 vulnerability such as poverty and inequality. Also, scientific knowledge is commonly prioritized over 12 traditional, Indigenous knowledge, and local knowledge. However, some transformative efforts are emerging 13 Bolivian Indigenous organizations provide a notable example by contesting normative conceptions of 14 development as economic growth with more comprehensive views like harmony with Mother Earth and 15 "Sumak Kawsay" or "Good Living". 16 17 Several strategies have been proposed to overcome existing barriers, including building effective and 18 respectful partnerships among Indigenous and non­indigenous researchers to co­produce climate change­ 19 relevant knowledge, and recognizing Indigenous Peoples as active actors who are continually developing 20 autonomous strategies to preserve their practices, beliefs and knowledge. The implementation of these and 21 other strategies can significantly enhance adaptation planning and action in the region. 22 23 [END FAQ 12.5 HERE] 24 25 26 Do Not Cite, Quote or Distribute 12-112 Total pages: 181