FINAL DRAFT                       Chapter 18               IPCC WGII Sixth Assessment Report

1

2                     Chapter 18: Climate Resilient Development Pathways

3

4 Coordinating Lead Authors: E. Lisa F. Schipper (Sweden/United Kingdom), Aromar Revi (India),
5 Benjamin L. Preston (Australia/USA)

 6

 7 Lead Authors: Edward. R. Carr (USA), Siri H. Eriksen (Norway), Luis R. Fern醤dez-Carril (Mexico),
 8 Bruce Glavovic (South Africa/New Zealand), Nathalie J.M. Hilmi (France/Monaco), Debora Ley
 9 (Mexico/Guatemala), Rupa Mukerji (India/Switzerland), M. Silvia Muylaert de Araujo (Brazil), Rosa Perez
10 (Philippines), Steven K. Rose (USA), Pramod K. Singh (India)

11

12 Contributing Authors: Paulina Aldunce (Chile), Aditya Bahadur (India), Nat醠ia Barbosa de Carvalho
13 (Brazil), Ritwika Basu (India), Donald A. Brown (USA), Anna Carthy (Ireland), Vanesa Cast醤 Broto
14 (Spain/United Kingdom), Ralph Chami (USA), John Cook (USA), Daniel de Berr阣o Viana (Brazil), Frode
15 Degvold (Norway), Shekoofeh Farahmend (Iran), Roger Few (United Kingdom), Gianfranco Gianfrate
16 (France), H. Carina Keskitalo (Sweden), Florian Krampe (Germany/Sweden), Rinchen Lama (South Africa),
17 Julia Leventon (Czech Republic/United Kingdom), Rebecca McNaught (New Zealand), Yu Mo
18 (China/United Kingdom), Marianne Mosberg (Norway), Michelle Mycoo (Trinidad and Tobago), Johanna
19 Nalau (Australia/Finland), Karen O'Brien (Norway), Meg Parsons (New Zealand), Alain Safa (France),
20 Zoha Shawoo (Pakistan/United Kingdom), Marcus Taylor (Canada/United Kingdom), Mark G.L. Tebboth
21 (United Kingdom), Bejoy K. Thomas (India), Kirsten Ulsrud (Norway), Saskia Werners (The
22 Netherlands/Germany), Keren Zhu (China), Monika Zurek (Germany/United Kingdom)

23

24 Review Editors: Diana Liverman (USA), Nobuo Mimura (Japan)

25

26 Chapter Scientists: Ritwika Basu (India/United Kingdom), Zoha Shawoo (Pakistan/United Kingdom), Yu
27 Mo (China/United Kingdom)

28

29 Date of Draft: 1 October 2021

30

31 Notes: TSU Compiled Version

32

33

34 Table of Contents

35

36 18.1 Ways Forward for Climate Resilient Development ..............................................................................8

37  18.1.1 Understanding Climate Resilient Development..............................................................................8

38  18.1.2 Pathways for Climate Resilient Development.................................................................................9

39  18.1.3 Policy Context for Climate Resilient Development .......................................................................13

40  18.1.4 Assessing Climate Resilient Development.....................................................................................14

41  18.1.5 Chapter Roadmap .........................................................................................................................15

42 Box 18.1: Transformations in Support of Climate Resilient Development Pathways .............................15

43 Box 18.2: Visions of Climate Resilient Development in Kenya ..................................................................17

44 18.2 Linking Development and Climate Action...........................................................................................18

45  18.2.1 Implications of Current Development Trends...............................................................................19

46  18.2.2 Understanding Development in Climate Resilient Development..................................................19

47  18.2.3 Scenarios as a Method for Representing Future Development Trajectories................................22

48  18.2.4 Climate Change Risks to Development.........................................................................................25

49  18.2.5 Options for Managing Future Risks to Climate Resilient Development.......................................26

50 Box 18.3: Climate Resilient Development in Small Islands ........................................................................30

51 Box 18.4: Adaptation and the Sustainable Development Goals .................................................................31

52 18.3 Transitions to Climate Resilient Development ....................................................................................42

53  18.3.1 System Transitions as a Foundation for Climate Resilient Development.....................................43

54 Box 18.5: The Implications of the Belt and Road Initiative (BRI) for Climate Resilient Development.45

55 Box 18.6: The Role of Ecosystems in Climate-Resilient Development ......................................................49

56 Cross-Chapter Box GENDER: Gender, Climate Justice and Transformative Pathways.......................57

57  18.3.2 Accelerating Transitions...............................................................................................................63

    Do Not Cite, Quote or Distribute  18-1                                Total pages: 197
    FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1 18.4 Agency and Empowerment for Climate Resilient Development........................................................64

2   18.4.1 Political Economy of Climate Resilient Development ..................................................................65

3   18.4.2 Enabling Conditions for Near-Term System Transitions..............................................................66

4 Box 18.7: `Green' Strategies of Institutional Investors...............................................................................70

5   18.4.3 Arenas of Engagement...................................................................................................................72

6 Cross-Chapter Box INDIG: The Role of Indigenous Knowledge and Local Knowledge in

7   Understanding and Adapting to Climate Change ...............................................................................74

8 Box 18.8: Macroeconomic policies in support of Climate-Resilient Development ...................................82

9   18.4.4 Frontiers of Climate Action ..........................................................................................................85

10 Box 18.9: The Role of the Private Sector in Climate Resilient Development via Climate Finance,

11  Investments and Innovation. .................................................................................................................86

12 18.5 Sectoral and Regional Synthesis of Climate Resilient Development .................................................86

13  18.5.1 Regional Synthesis of Climate-Resilient Development .................................................................86

14  18.5.2 Sectoral Synthesis of Climate-Resilient Development ..................................................................91

15  18.5.3 Feasibility and Efficacy of Options for Climate-Resilient Development......................................92

16 18.6 Conclusions and Research Needs ........................................................................................................104

17  18.6.1 Knowledge Gaps .........................................................................................................................104

18  18.6.2 Conclusions.................................................................................................................................105

19 FAQ18.1: What is a climate resilient development pathway?..................................................................106

20 FAQ18.2: What is climate resilient development and how can climate change adaptation (measures)

21  contribute to achieving this?................................................................................................................106

22 FAQ18.3: How can different actors across society and levels of government be empowered to pursue

23  climate resilient development? ............................................................................................................108

24 FAQ18.4: What role do transitions and transformations in energy, urban and infrastructure,

25  industrial, land and ocean ecosystems, and in society, play in climate resilient development? ....108

26 FAQ18.5: What are success criteria in climate resilient development and how can actors satisfy those

27  criteria? .................................................................................................................................................109

28 References......................................................................................................................................................110

29

30 Cross-Chapter Box FEASIB: Feasibility Assessment of Adaptation Options: An Update of the SR1.5

31  ................................................................................................................................................................ 154

32

33

    Do Not Cite, Quote or Distribute  18-2                                          Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1 Executive Summary

 2 Climate resilient development (CRD) is a process of implementing greenhouse gas mitigation and
 3 adaptation options to support sustainable development for all (18.1). Climate action and sustainable
 4 development are interdependent processes and climate resilient development is possible when this
 5 interdependence is leveraged. Pursuing these goals in an integrated manner increases their effectiveness in
 6 enhancing human and ecological well-being. Climate resilient development can help build capacity for
 7 climate action, including contributing to reductions in greenhouse gas emissions while enabling the
 8 implementation of adaptation options that enhance social, economic and ecological resilience to climate
 9 change as the prospect of crossing the 1.5癈 global warming level in the early 2030s approaches (WG1
10 Table SPM1). For example, incorporating clean energy generation, healthy diets from sustainable food
11 systems, appropriate urban planning and transport, universal health coverage and social protection, can
12 generate substantial health and wellbeing co-benefits (very high confidence1) (7.4.4, Cross-Chapter Box
13 HEALTH in Chapter 7). Similarly, universal water and energy access can help to reduce poverty and
14 improve well-being while making populations less vulnerable and more resilient to adverse climate impacts
15 (very high confidence) (18.1, Box 4.7).

16

17 Current development pathways combined with the observed impacts of climate change, are leading
18 away from, rather than toward, sustainable development, as reported in recent literature (moderate
19 agreement, robust evidence). While demonstrable progress has been made on some of the SDGs, significant
20 gains across a range of targets are still necessary, as is enhancing synergies and balancing and managing
21 trade-offs. Severe risks to natural and human systems are already observed in some places (high confidence),
22 and could occur in many more systems, worldwide before mid-century (medium confidence), by end-century
23 at all scales, from the local to the global, and at all latitudes and altitudes (high confidence). The COVID-19
24 pandemic revealed the vulnerability of development progress to shocks and stresses, potentially delaying the
25 implementation of the 2030 Agenda for all (8.1, Cross-Chapter Box COVID in Chapter 7). Various global
26 trends including rising income inequality, continued growth in greenhouse gas emissions, land use change,
27 food and water insecurity, human displacement, and reversals of long-term increasing life expectancy trends
28 in some nations run counter to the SDGs (very high confidence) as well as efforts to mitigate greenhouse gas
29 emissions and adapt to a changing climate (18.2). These development trends contribute to worsening
30 poverty, injustice and inequity, and environmental degradation. Climate change can exacerbate these
31 conditions by undermining human and ecological well-being (18.2).

32

33 Social and economic inequities linked to gender, poverty, race/ethnicity, religion, age, or geographic
34 location compound vulnerability to climate change and have created and could further exacerbate
35 injustices, and constrain the implementation of CRD for all (very high confidence). Climate change
36 intensifies existing vulnerability and inequality, with adverse impacts of climate change on the most
37 vulnerable groups, including women and children in low-income households, Indigenous or other minority
38 groups, small-scale producers and fishing communities, and low-income countries (high confidence). Most
39 vulnerable regions and population groups, such as in East, Central and West Africa, South Asia, Micronesia
40 and Melanesia and in Central America, present the most urgent need for adaptation (high confidence) (Ch
41 10, 12, 15). Climate justice initiatives explicitly address these multi-dimensional distributional issues as part
42 of climate change adaptation. However, adaptation strategies can worsen social inequities, including gender,
43 unless explicit efforts are made to change those unequal power dynamics, including spaces to foster
44 inclusive decision-making. Drawing upon Indigenous knowledge and local knowledge can contribute to
45 overcoming the combined challenges of climate change, food security, biodiversity conservation, and
46 combating desertification and land degradation. (18.2; Cross-Chapter Box GENDER; Cross-Chapter Box
47 INDIG}

48

49 Opportunities for climate resilient development vary by location (very high confidence). Over 3.3
50 billion people live in regions that are very high and highly vulnerable to climate change, while 2 billion
51 people live in regions with low and very low vulnerability. Response to global greenhouse gas emissions

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  18-3        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 trajectories, regional and local development pathways, climate risk exposure, socio-economic and ecological
 2 vulnerability, and the local capacity to implement effective adaptation and greenhouse gas mitigation
 3 options, differ depending on local contexts and conditions (Table 18.3). As an example, underlying social
 4 and economic vulnerabilities in Australasia, exacerbate disadvantage among particular social groups and
 5 there is deep underinvestment in adaptation, given current and projected risks (Ch 11). There is also
 6 significant regional heterogeneity in climate change, exposure, and vulnerability, indicating different starting
 7 points for CRD, as well as mitigation, adaptation, and sustainable development opportunities, synergies, and
 8 trade-offs (18.5).

 9

10 There are multiple possible pathways by which communities, nations and the world can pursue
11 climate resilient development. Moving toward different pathways involves confronting complex
12 synergies and trade-offs between development pathways, and the options, contested values, and
13 interests that underpin climate mitigation and adaptation choices (very high confidence). Climate
14 resilient development pathways are trajectories for the pursuit of climate resilient development and
15 navigating its complexities. Different actors, the private sector, and civil society, influenced by science, local
16 and Indigenous knowledges, and the media are both active and passive in designing and navigating CRD
17 pathways (18.1, 18.4). Increasing levels of warming may narrow the options and choices available for local
18 survival and sustainable development for human societies and ecosystems. Limiting warming to Paris
19 Agreement goals will reduce the magnitude of climate risks to which people, places, the economy and
20 ecosystems will have to adapt. Reconciling the costs, benefits, and trade-offs associated with adaptation,
21 mitigation, and sustainable development interventions and how they are distributed among different
22 populations and geographies is essential and challenging, but also creates the potential to pursue synergies
23 that benefit human and ecological well-being. For example, in parts of Asia sustainable development
24 pathways that connect climate change adaptation and disaster risk reduction can reduce climate vulnerability
25 and increase resilience (Table 18.3. 10.6.2). Different actors and stakeholders have different priorities
26 regarding these opportunities, which can exacerbate or diminish existing social, economic and ecological
27 vulnerabilities and inequities. For example, in parts of Africa, intensive irrigation contributes to the
28 development of agriculture but has come at a cost to ecosystem integrity and human well-being (Table 18.3.,
29 9.15.2). Careful and explicit consideration for the ethical and equity dimensions of policies and practices
30 associated with a climate resilient development pathway can help limit these negative externalities.

31

32 Prevailing development pathways are not advancing climate resilient development (very high
33 confidence). Societal choices in the near-term will determine future pathways. Some low-emissions
34 pathways and climate outcomes are unlikely2 to be realized (very high confidence). Rapid climate change is
35 affecting every region across the globe and affecting natural and human systems relevant to the pursuit of the
36 SDGs (18.1, 18.2, Fig. 18.1). Even the most ambitious greenhouse gas mitigation scenarios indicate climate
37 change will continue for decades to centuries (WGI, 18.2). Increasing mitigation effort across multiple
38 sectors exhibits opportunities for synergies with sustainable development, but also trade-offs that increase
39 with mitigation effort that need to be balanced and managed (high confidence). The uncertainty associated
40 with achieving specific pathways and climate outcomes is a risk factor to consider in planning, with
41 plausibility and transformational challenges, as well as trade-offs and synergies, affected by technology,
42 policy design, and societal choices (18.2). For instance, restrictions on utilization of individual mitigation
43 options to manage trade-offs (e.g., bioenergy with CCS, afforestation, nuclear power) can also affect the
44 mitigation cost to households (e.g., energy security, commodity prices) and the likelihood of a desired
45 climate outcome being realized. Developing and transitional economies are estimated as low-cost mitigation
46 opportunities, but are often at high risk from climate change due to their regional and development context
47 (high confidence) (18.2,18.5). For example in Africa, competing uses for water such as hydropower
48 generation, irrigation, and ecosystem requirements can create trade-offs among different management and
49 development objectives (9.7.3). In Asia, intensive irrigation and other forms of water consumption can
50 have a negative effect on water quality and aquatic ecosystems (Ch 10.6.3). Developed countries also,

2

 In this Report, the following terms have been used to indicate the assessed likelihood of an outcome or a result:
Virtually certain 99�100% probability, Very likely 90�100%, Likely 66�100%, About as likely as not 33�66%,
Unlikely 0�33%, Very unlikely 0�10%, and Exceptionally unlikely 0�1%. Additional terms (Extremely likely: 95�
100%, More likely than not >50�100%, and Extremely unlikely 0�5%) 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  18-4        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 face trade-offs, including in Australasia where adapting to fire risk in peri-urban zones introduces potential
 2 trade-offs among ecological values and fuel reduction in treed landscapes (Ch 11.3.5) and in North America
 3 where new coastal and alpine developments generate economic activity but enhance local social inequalities
 4 (15.4.10).

 5

 6 Systems transitions can enable climate resilient development, when accompanied by appropriate
 7 enabling conditions and inclusive arenas of engagement (very high confidence). Five systems transitions
 8 are considered: energy, industry, urban and infrastructure, land and ecosystems, and societal. Advancing
 9 climate resilient development in specific contexts may necessitate simultaneous progress on all five
10 transitions. Collectively, these system transitions can widen the solution space and accelerate and deepen the
11 implementation of sustainable development, adaptation, and mitigation actions by equipping actors and
12 decision-makers with more effective options. For example, urban ecological infrastructure linked to an
13 appropriate land use mix, street connectivity, open and green spaces, and job-housing proximity provides
14 adaptation and mitigation benefits that can aid urban transformation. (Table 18.4, Cross-Working Group Box
15 URBAN in Chapter 6) These system transitions are necessary precursors for more fundamental climate and
16 sustainable-development transformations; but can simultaneously be outcomes of transformative actions.
17 However, the way they are pursued may not necessarily be perceived as ethical or desirable to all actors.
18 Hence, enhancing equity and agency are cross-cutting considerations for all five transitions. Such transitions
19 can generate benefits across different sectors and regions, provided they are facilitated by appropriate
20 enabling conditions including effective governance, policy implementation, innovation, and climate and
21 development finance, which are currently insufficient (18.3, 18.4).

22

23 There is a rapidly narrowing window of opportunity to implement system transitions needed to enable
24 CRD. Past choices have already eliminated some development pathways, but other pathways for
25 climate-resilient development remain (very high confidence). In spite of a growth in national net-zero
26 commitments, the current prospects of surpassing 1.5癈 global mean temperatures by the 2030s are high

27 (WG1 Table SPM1). There is strong evidence of the worsening of multiple climate impact drivers

28 in all regions, that will place additional pressures on ecosystem services that support food and water

29 systems, increasing the risks of malnutrition, ill-health and poverty in many regions (WG1 Fig

30 SPM9, Table 18.4). This implies that significant additional adaptation will be needed. Over the

31 near-term, implementing such transformational change could be disruptive to various economic and

32 social systems. Over the long-term, however, they could generate benefits to human well-being and

33 planetary health. Strengthening coordinated adaptation and mitigation actions can enhance the

34 potential of local and regional development pathways to support CRD. Planning for CRD can

35 support both adaptation and decarbonization via effective land-use, promoting resilient and low-

36 carbon infrastructure; protecting biodiversity and integrating ecosystem services (Table 18.4),

37 assuming advancing just and equitable development processes.

38

39 Prospects for transformation towards climate resilient development increase when key governance
40 actors work together in inclusive and constructive ways to create a set of appropriate enabling
41 conditions (18.4.2) (high confidence). These enabling conditions include effective governance and
42 information flow, policy frameworks that incentivize sustainability solutions; adequate financing for
43 adaptation, mitigation, and sustainable development; institutional capacity; science, technology and
44 innovation; monitoring and evaluation of climate resilient development policies, programs, and practices;
45 and international cooperation. Investment in social and technological innovation, could generate the
46 knowledge and entrepreneurship needed to catalyze system transitions, and their transfer. The
47 implementation of policies that incentivize the deployment of low-carbon technologies and practices within
48 specific sectors such as energy, buildings, and agriculture could accelerate greenhouse gas mitigation and
49 deployment of climate resilient infrastructure, in urban and rural areas. Civic engagement is an important
50 element of building societal consensus and reducing barriers to action on adaptation, mitigation, and
51 sustainable development. (18.4)

52

53 CRD pathways are determined through engagement in different arenas degree to which the emergent
54 pathways foster just, and climate resilient development depends on how contending societal interests,
55 values and worldviews are reconciled through inclusive and participatory interactions between
56 governance actors in these arenas of engagement (18.4.3) (high confidence). These interactions occur in

Do Not Cite, Quote or Distribute  18-5        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 many different arenas (e.g., governmental, economic and financial, political, knowledge, science &
 2 technology, and community) that represent the settings, places, and spaces in which societal actors interact to
 3 influence the nature and course of development. For instance, the Agenda 2030 highlights the importance of
 4 multi-level adaptation governance, including non-state actors from civil society and the private sector. This
 5 implies the need for wider arenas and modes of engagement around adaptation that facilitate coordination,
 6 convergence, and productive contestation among these diverse actors to collectively solve problems and to
 7 unlock the synergies between adaptation and mitigation and sustainable development.

 8

 9 Regional and national differences mean different capacities for pursuing climate resilient development
10 pathways. Economic sectors and global regions are exposed to different opportunities and challenges
11 in facilitating climate resilient development, suggesting adaptation and mitigation options should be
12 aligned to local and regional context and development pathways (very high confidence). Given their
13 current state of development, some regions may prioritize poverty and inequality reduction, and economic
14 development over the near-term as a means of building capacity for climate action and low-carbon
15 development over the long-term. For example, Africa, South Asia, and Central and South America are highly
16 exposed, vulnerable and impacted by climate change, which is amplified by poverty, population growth, land
17 use change and high dependence on natural resources for commodity production. In contrast, developed
18 economies with mature economies and high levels of resilience may prioritize climate action to transition
19 their energy systems and reduce greenhouse gas emissions. Some interventions may be robust in that they
20 are relevant to a broad range of potential development trajectories and could be deployed in a flexible
21 manner. For example, conservation of land and water could be achieved through a variety of means and offer
22 benefits to populations in the global North and South alike. However, other types of interventions, such as
23 those that are dependent upon emerging technologies, may require a specific set of enhanced enabling
24 conditions or factors including infrastructure, supply chains, international cooperation, and education and
25 training that currently limit their implementation to certain settings (18.5). Notwithstanding national and
26 regional differences, development practices that are aligned to people, prosperity, partnerships, peace and the
27 planet as defined in Agenda 2030, could enable more climate resilient development (see Figure 18.1).

28

29 People, acting through enabling social, economic and political institutions, are the agents of system
30 transitions and societal transformations that facilitate climate resilient development founded on the
31 principles of inclusion, equity, climate justice, ecosystem health, and human well-being (very high
32 confidence). While much literature on climate action has focused on the role of technology and policy as the
33 factors that drive change, recent literature has focused on the role of specific actors � citizens, civil society,
34 knowledge institutions (including local and Indigenous Peoples and science), governments, investors and
35 businesses. Greater attention to, and transparency of, which actors' benefit, fail to benefit, or are impacted by
36 mitigation and adaptation choices actions could better support climate-resilient and sustainable development.
37 For example, grounding adaptation actions in local realities could help to ensure that adaptive actions do not
38 worsen existing gender and other inequities within society (e.g., leading to maladaptation practices) (high
39 confidence). Differences in the ability of different actors to effect change ultimately influence which
40 interventions for sustainable development or climate action are implemented and thus what development
41 outcomes are achieved. Recent literature has focused on the social, political, and economic arenas of
42 engagement, in which these different actors interact. More focused attention on these arenas of engagement
43 could prove beneficial to reconciling divergent views on climate action, integrating Indigenous knowledge
44 and local knowledges, elevating diverse voices that have historically been marginalized from the policy
45 discourse, thereby reducing vulnerability, deepening adaptive capacity and the ability to implement CRD
46 (18.4; Cross-Chapter Box GENDER; Cross-Chapter Box INDIG)

47

48 Pursuing climate resilient development involves considering a broader range of sustainable
49 development priorities, policies and practices, as well as enabling societal choices to accelerate and
50 deepen their implementation (very high confidence). Scientific assessments of climate change have
51 traditionally framed solutions around the implementation of specific adaptation and mitigation options as
52 mechanisms for reducing climate-related risks. They have given less attention to a fuller set of societal
53 priorities and the role of non-climate policies, social norms, lifestyles, power relationships and worldviews in
54 enabling climate action and sustainable development. Because climate resilient development involves
55 different actors pursuing plural development trajectories in diverse contexts, the pursuit of solutions that are
56 equitable for all requires opening the space for engagement and action to a diversity of people, institutions,
57 forms of knowledge, and worldviews. Through inclusive modes of engagement that enhance knowledge

Do Not Cite, Quote or Distribute  18-6        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1 sharing and realize the productive potential of diverse perspectives and worldviews, societies could alter
2 institutional structures and arrangements, development processes, choices and actions that have precipitated
3 dangerous climate change, constrained the achievement of SDGs, and thus limited pathways to achieving
4 CRD (Box 18.1, 18.4). There are only a few decades remaining to chart CRD pathways that catalyze the
5 transformation of prevailing development practices and offer the greatest promise and potential for human
6 well-being and planetary health.

7

Do Not Cite, Quote or Distribute  18-7        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 18.1 Ways Forward for Climate Resilient Development

 2

 3 The links between climate change and development have been long recognized by various research
 4 communities (Nagoda, 2015; Winkler et al., 2015; Webber, 2016; Carr, 2019) and have been assessed by
 5 Working Group II in every IPCC Assessment Report since AR3 (Smit et al., 2001; Yohe et al., 2007; Denton
 6 et al., 2014). For the AR1-3 reports, these links were largely framed in the context of sustainable
 7 development, a concept that has been well described in the literature for decades (Brundtland, 1987). The
 8 AR5 introduced the framing of climate resilient pathways, which narrowed the discussion around sustainable
 9 development to specifically address the contributions of mitigation and adaptation actions to the reduction of
10 risk to development and the various institutions, strategies, and choices involved in risk management
11 (Denton et al., 2014). That assessment concluded that identifying and implementing appropriate technical
12 and governance options for mitigation and adaptation as well as development strategies and choices that
13 contribute to climate resilience are central to the successful implementation of such strategies. The AR5 also
14 recognized that transformation of current development pathways in terms of wider political, economic and
15 social systems may be necessary (Denton et al., 2014).

16

17 The literature presenting research findings on climate resilient development (CRD) and pathways and
18 processes for successfully achieving CRD has expanded significantly in the several years since the AR5
19 (very high confidence). This includes both qualitative studies of development as well as illustrative,
20 quantitative analyses of development trajectories linked to specific scenarios, such as the Shared
21 Socioeconomic Pathways (SSPs) (18.2.2). Furthermore, the literature describing the role of system
22 transitions and societal transformation in enabling climate action (Box 18.1, 18.3), compliance with the Paris
23 Agreement (18.1.3, 18.2.1), and achievement of the Sustainable Development Goals (18.1.3; Box 18.4) has
24 expanded significantly (very high confidence). This expansion is comprised of studies spanning a broad
25 range of disciplinary perspectives, some of which have been underrepresented in prior IPCC assessments
26 (high agreement, limited evidence) (Minx et al., 2017; Pearce et al., 2018b)).

27

28 This chapter therefore focuses on assessing this more recent literature and the diverse scientific
29 understandings of CRD and the pathways for pursuing it. Notably, this chapter takes off where Chapters 16
30 and 17 end: recognizing the decision-making context to address the representative key risks and their
31 intersections with development, among others. This chapter therefore highlights not only how climate risk
32 undermines CRD, but also how current patterns of development contribute to climate risk, both generally
33 and in different sectoral and regional contexts. In particular, the chapter focuses on achieving CRD through
34 systems transitions, discussing these in relation to societal transformation, and how different actors engage
35 one another in order to pursue policy and practice consistent with CRD.

36

37 18.1.1 Understanding Climate Resilient Development

38

39 Past IPCC Assessment Reports have consistently examined an extensive literature on the links between
40 climate change, adaptation, and sustainable development (Smit et al., 2001; Klein et al., 2007; Yohe et al.,
41 2007). However, studies that explicitly refer to CRD as a concept or a guide for policy and practice remain
42 modest (very high confidence). The concept of CRD appeared in scholarly literature as well as development
43 program documents over a decade ago (Kamal Uddin et al., 2006; Garg and Halsn鎠, 2007) and has been
44 used in more recent IPCC assessment reports and special reports (e.g., Denton et al., 2014; Roy et al., 2018).
45 Similarly, the use of the term climate resilient development pathways dates to 2009 (Ayers and Huq, 2009),
46 but its use accelerated after appearing in UNFCCC publications around the launch of the Green Climate
47 Fund (UNFCCC, 2011). While this chapter prioritizes the CRD literature, it also recognizes a broad range of
48 literature, disciplinary expertise, and development practice is relevant to the concept of CRD.

49

50 Much of this literature is assessed in recent IPCC Special Reports (Rogelj et al., 2018; Roy et al., 2018;
51 Bindoff et al., 2019; Hurlbert et al., 2019; Oppenheimer et al., 2019), but new studies have continued to
52 emerge. More specific uses of CRD found in the literature describe development that seeks to achieve
53 poverty reduction and adaptation to climate change simultaneously without explicit mention of mitigation
54 (USAID, 2014)), as well as mitigation and poverty reduction, described as `low-carbon development,'
55 without explicit mention of adaptation (Alam et al., 2011; Fankhauser and McDermott, 2016). Other similar
56 terms include `climate safe', `climate compatible' and `climate smart' development (Huxham et al., 2015;
57 Kim et al., 2017b; Ficklin et al., 2018; Mcleod et al., 2018), each with varying nuances. Climate-compatible

Do Not Cite, Quote or Distribute  18-8        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 development coined by Mitchell and Maxwell (2010) specifically describes a `triple win' of adaptation,
 2 mitigation and development (Antwi-Agyei et al., 2017; Favretto et al., 2018) (see also 8.6). In this spirit,
 3 AR5 specifically referred to climate-resilient development as "development trajectories that combine
 4 adaptation and mitigation to realize the goal of sustainable development" (Denton et al., 2014). This chapter
 5 builds on the AR5 and, for the purposes of assessment, formally defines CRD as a process of implementing
 6 greenhouse gas mitigation and adaptation measures to support sustainable development for all. This
 7 extension of the earlier definition reflects the emphasis in recent literature on equity as a core element of
 8 sustainable development as well as the objective of the SDGs to "create conditions for sustainable, inclusive
 9 and sustained economic growth, shared prosperity and decent work for all, taking into account different
10 levels of national development and capacities" (United Nations, 2015: 3/35).

11

12 Past, present, and future concentrations of greenhouse gases in the atmosphere are the direct result of both
13 natural and anthropogenic greenhouse gas emissions which are, in turn, a function of past and current
14 patterns of human and economic development (very high confidence, WGI SPM). This includes development
15 processes that drive land use change, extractive industries, manufacturing and trade, energy production, food
16 production, infrastructure development, and transportation. These patterns of development are therefore
17 drivers of current and future climate risk to specific sectors, regions, and populations (Byers et al., 2018), as
18 well as the demand for both mitigation and adaptation as a means of preventing climate change from
19 undermining development goals. The Sustainable Development Goals (SDGs) represent targets for
20 supporting human and ecological well-being in a sustainable manner. Yet, while progress is being made
21 toward a number of the Sustainable Development Goals (SDGs), success in achieving all of the SDGs by
22 2030 across all global regions remains uncertain (high agreement, medium evidence) (United Nations, 2021).
23 Moreover, current commitments to reduce greenhouse gas emissions are not yet consistent with limiting
24 changes in global mean temperature elevation to less than 2癈 or 1.5癈 (very high confidence) (IPCC,
25 2018a) (see also 18.2).

26

27 Atmospheric concentrations of greenhouse gases are just one of a number of planetary boundaries which
28 define safe operating spaces for humanity and therefore opportunities for achieving sustainable and climate-
29 resilient development. Exceeding these boundaries poses increased risk of large-scale abrupt or irreversible
30 environmental changes that would threaten human and ecological well-being (very high confidence)
31 (Rockstr鰉 et al., 2009a; Rockstr鰉 et al., 2009b; Butler, 2017; Schleussner et al., 2021). Other planetary
32 boundaries reported in the literature such as biodiversity loss, changes in land systems, and freshwater use
33 are also directly influenced by patterns of development as well as climate change (18.2; 18.5). Current rates
34 of species extinction, conversion of land for crop production, and exploitation of water resources exceed
35 planetary boundaries, thereby undermining CRD. Moreover, studies indicate that achievement of the
36 sustainable development goals, while consistent with maintaining some planetary boundaries, could
37 undermine others (O'Neill et al., 2018; Hickel, 2019; Randers et al., 2019) (18.2), suggesting significant
38 shifts in current patterns of development are necessary to maintain development within planetary boundaries.

39

40 Exceedance of planetary boundaries contributes to human and ecological vulnerability to climate change and
41 other shocks and stressors. People and regions that already face high rates of natural resource use, ecosystem
42 degradation, and poverty are more vulnerable to climate change impacts, compounding existing development
43 challenges in regions that are already strained (IPCC, 2014a; Hallegatte et al., 2019). The International
44 Monetary Fund, for example, found that for a medium and low-income developing country with an annual
45 average temperature of 25癈, the effect of a 1癈 increase in temperature is a reduction in economic growth
46 by 1.2% (Acevedo et al., 2018). Countries whose economies are projected to be hard hit by an increase in
47 temperature account for only about 20% of global Gross Domestic Product (GDP) in 2016, but are home to
48 nearly 60% of the global population. This is expected to rise to more than 75% by the end of the century.
49 These economic impacts are a function of the underlying vulnerability of low- and middle-income
50 developing economies to the impacts of climate change (see 18.5). Such vulnerability was also evidenced
51 and enhanced by the COVID-19 pandemic which slowed progress on the SDGs in multiple nations (Naidoo
52 and Fisher, 2020; Srivastava et al., 2020; Bherwani et al., 2021).

53

54 18.1.2 Pathways for Climate Resilient Development

55

56 One approach for operationalizing the concept of climate-resilient development in a decision-making context
57 is to link the concept of CRD to that of pathways (Figure 18.1). A pathway can be defined as "a trajectory in

Do Not Cite, Quote or Distribute  18-9        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 time, reflecting a particular sequence of actions and consequences against a background of autonomous
 2 developments, leading to a specific future situation" (Haasnoot et al., 2013; Bourgeois, 2015). As such, a
 3 pathway represents changes over time in response to policies and practices as well spontaneous and
 4 exogenous events. For example, the SR1.5 report suggested that CRD pathways are "a conceptual and
 5 aspirational idea for steering societies towards low-carbon, prosperous and ecologically safe futures" (Roy
 6 et al., 2018: 468), and a way to highlight the complexity of decision-making processes at different levels.
 7 Here, consistent with the aforementioned definition of CRD, we define CRD pathways as development
 8 trajectories that successfully integrate mitigation, adaptation, and sustainable development to achieve
 9 development goals.

10

11 As illustrated in Figure 18.1, the ultimate aim of CRD pathways is to support sustainable development for
12 ensuring planetary health and human well-being. CRD is both an outcome at a point in space and time, as
13 observed through SDG achievement indicators, but also a process consisting of actions and social choices
14 made by multiple actors--government, industry, media, civil society, and science (18.4). These actions and
15 social choices are performed within different dimensions of governance--politics, institutions (norms, rules),
16 and practice, and bounded by ethics, values and worldviews. The development outcomes and processes
17 pertain to political, economic, ecological, socio-cultural, knowledge-technology, and community arenas
18 (Figure 18.2). A CRDP will, for example, aspire to achieve ecological outcomes in terms of planetary health
19 and achievement of Paris Agreement goals as well as human well-being, solidarity and social justice, in
20 addition to political, economic, and science-technology outcomes. These outcomes are enabled by achieving
21 progress in core system transitions that catalyze broader societal transformations (Figure 18.3).

22

23 While there are many possible successful pathways to future development in the context of climate change,
24 history has shown that pathways that are positive for the vast majority, often induce notable impacts and
25 costs, especially on marginal and vulnerable people (Hickel, 2017; Ramalho, 2019), placing them in direct
26 contradiction with the commitment to `leave no one behind' (United Nations, 2015). Similarly,
27 contemporary scenario analyses find that there are plausible development trajectories that lead toward
28 sustainability (Figure 18.1, 18.2.2). Yet, a number of plausible trajectories that perpetuate or exacerbate
29 unstainable forms of development also appear in the literature (Figure 18.1, 18.2.2). A significant challenge
30 lies in identifying pathways that address current climate variability and change, while allowing for
31 improvements in human well-being. Furthermore, while a given pathway might lead to a set of desired
32 outcomes for one region or set of actors, the process of getting there may come at high environmental, socio-
33 and economic cost to others (very high confidence) (Raworth, 2017; Faist, 2018). Frequently, considerations
34 of social difference and equity are not prioritized in the evaluation of different development choices. The
35 assumption that a growing economy lifts opportunity for all, could for example, further marginalize those
36 who are the most vulnerable to climate change (Matin et al., 2018; Diffenbaugh and Burke, 2019; Hickel et
37 al., 2021).

38

39 Placing pathways and climate actions within development processes implies a broadening of enablers to
40 include the ethical-political quality of socio-environmental processes that are required to shift such processes
41 in directions that support CRD and the pursuit of sustainability outcomes. This chapter therefore departs
42 from the AR5s alignment of CRD with adaptation pathways and the emphasis on decision points that enable
43 one to manage (or fail to manage) climate risk towards a framing that integrates a range of possible futures
44 each offering different opportunities, risks, and trade-offs to different actors and stakeholders (see WGII
45 AR5, IPCC, 2014b, Figure SPM.9). Instead, CRD emerges from everyday formal and informal decisions,
46 actions, and adaptation or mitigation policy interventions. This is inclusive of system transitions, increased
47 resilience, environmental integrity, social justice, equity, and reduced poverty and vulnerability, all facets of
48 human well-being and planetary health. Rather than encompassing a formula or blueprint for particular
49 actions, sustainable development is a process that provides a compass for the direction that these multiple
50 actions should take (Anders, 2016). This creates opportunities for actors to apply a diverse toolkit of
51 adaptation, mitigation, and sustainable development interventions, thereby opening up the solution space.

52

53

Do Not Cite, Quote or Distribute  18-10       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1

 2 Figure 18.1: Climate Resilient Development Pathways. Climate resilient development is a process that takes place
 3 through societal choices towards (green pathways) or away from (red pathways) five development dimensions (people,
 4 prosperity, partnership, peace, planet) on which the SDGs build. Some societal choices have mixed outcomes for CRD
 5 (orange pathways). This figure builds on figure SPM.9 in AR5 WGII depicting climate resilient pathways) by
 6 describing how CRDPs emerge from societal choices within multiple arenas � rather than solely from discrete decision
 7 points. Societal choices, often contested, are made in these arenas through interactions between key actors in civil
 8 society, the private sector and government (see Figure 18.2). The quality of interactions between these actors in these
 9 arenas determine whether societal choices shift development towards or away from CRD. For example, inclusion vs
10 exclusion and influence over choices shapes the quality of these interactions, and the outcomes of emergent societal
11 choices. These qualities thus also characterize alternative futures resulting from different pathways, along five
12 development dimensions (people, prosperity, partnership, peace, planet) on which the SDGs build. five CRD
13 dimensions underline the close interconnectedness between the biosphere and humans, the two necessarily intertwined
14 in interactions, actions, transitions, and futures (Figure 18.3). There is a narrow and closing window of opportunity to
15 make transformational changes to move towards and not away from development futures that are more climate-resilient
16 and sustainable. Pathways not taken (dotted line) show that the pathways towards the highest CRD futures are no longer
17 available due to past societal choices and increasing temperatures. Present societal choices determine whether we shift
18 towards CRD in future or whether pathways will be limited to less CRD.

Do Not Cite, Quote or Distribute  18-11       Total pages: 197
      FINAL DRAFT                 Chapter 18  IPCC WGII Sixth Assessment Report

1
2

 3

 4 Figure 18.2: Societal choices in arenas of engagement shaping actions and systems. The settings, places and spaces
 5 in which key actors from government, civil society and the private sector interact to influence the nature and course of
 6 development can be called arenas of engagement, including political, economic, socio-cultural, ecological, knowledge-
 7 technology and community arenas. For instance, political arenas include formal political settings such as voting
 8 procedures to elect local representatives as well as less formal and transparent political arenas. Streets, town squares
 9 and post-disaster landscapes can become sites of interaction and political struggle as citizens strive to have their voices
10 heard. Arenas exist across scales from the local to national level, and beyond. Arenas of engagement can take the form
11 of "struggle arenas" � in which power and influence are used to include/exclude, set agendas, and make and implement
12 decisions � with inevitable winners and losers. The quality of interactions in these arenas leads to development
13 outcomes that can be characterized as CRD dimensions that underpin the SDGs � people, prosperity, partnership, peace,
14 planet (see Figure 18.1). a) Interactions characterized by inequitable relations and domination of some actors over
15 others may lead to societal choices away from CRD, including exacerbating disempowerment and vulnerability among
16 marginalized groups. b) Prospects for moving towards CRD increase when governance actors work together
17 constructively in these different arenas. Interactions and actions that are inclusive and synchronous, as opposed to
18 fragmented or contradictory, enable system transitions and transformational change towards CRD (Figure 18.3b, Box
19 18.3). b) Well-intentioned efforts often fail to be transformative, but instead entrench inequities. Instead, marginalized
20 groups and future trends in vulnerability need to be placed at the center of efforts to chart CRDPs. Unlocking the
21 productive potential of conflict that often characterizes interactions in these arenas of engagement is central to
22 advancing human well-being and planetary health. Moreover, the window for doing so is closing rapidly to avert
23 dangerous climate change and unsustainable development.

24

25

26

27 Figure 18.3: Transformative actions and system transitions a) Societal choices that generate fragmented climate action
28 or inaction and unsustainable development perpetuate business as usual development. b) Societal choices that support
29 CRD involve transformative actions that drive five systems transitions (energy, land and other ecosystems, urban and
30 infrastructure, industrial and societal). There is close interdependence between these systems. The system transition

Do Not Cite, Quote or Distribute  18-12       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 framework allows for a comprehensive assessment of the synergies and trade-offs between mitigation, adaptation and
 2 sustainable development. For example, land and water use in one system impacts the other systems and their
 3 surrounding ecosystems, thus reflecting how agricultural practices can have an impact on energy usage in urban centers.
 4 Finally, societal system transitions within each of the other systems enable the transitions to occur

 5

 6

 7 This understanding of CRD implies that different actors � governments, businesses, and civic organizations
 8 � will have to design and navigate their own CRD pathways toward climate resilient and sustainable
 9 development. This includes determining the appropriate balance of adaptation, mitigation, and sustainable
10 development actions and investments that are consistent with individual actors' development circumstances
11 and goals while also ensuring that the collective actions remain consistent with global agreements and goals
12 (such as the SDGs, Sendai Framework, and the Paris Agreement; 18.1.3), planetary boundaries, and other
13 principles of CRD including social justice and equity (Roy et al., 2018). Empowering individual actors to
14 pursue CRD in context-specific manner while coordinating action among actors and a diversity of scales,
15 local to global, is a key challenge associated with achieving CRD (high agreement, limited evidence).

16

17 18.1.3 Policy Context for Climate Resilient Development

18

19 As reflected in Chapter 1 of the AR6 WGII report, CRD is emerging as one of the guiding principles for
20 climate policy, both at the international level (Denton et al., 2014; Segger, 2016), as reflected in the Paris
21 Agreement (Article 2, UNFCCC, 2015), and within specific countries (Simonet and Jobbins, 2016; Kim et
22 al., 2017b; Vincent and Colenbrander, 2018; Yalew, 2020). This framing of development recognizes the
23 risks posed by climate change to development objectives (18.2; see also Chapter 16); the opportunities,
24 constraints and limits associated with reducing risk through adaptation; synergies and trade-offs between
25 mitigation, adaptation, and sustainable development (18.2.5, 18.5, Box 18.4); and the role of system
26 transitions in enabling large-scale transformations that limit future global warming to less than 1.5癈 while
27 boosting resilience (IPCC, 2018a) (18.3, Box 18.1).

28

29 Since the AR5, the volume of research at the nexus of climate action and sustainable development has
30 changed markedly (very high confidence). A rapidly growing, multi-disciplinary literature has emerged on
31 climate resilient development (Mitchell et al., 2015; Clapp and Sillmann, 2019; Hardoy et al., 2019; Yalew,
32 2020) and associated pathways (Naess et al., 2015; Winkler and Dubash, 2016; Brechin and Espinoza, 2017;
33 Solecki et al., 2017; Ellis and Tschakert, 2019) (18.2.2). Nevertheless, the concept of resilience generally,
34 and climate resilient development specifically, has come under increasing criticism in recent years (very high
35 confidence) (Joakim et al., 2015; Schlosberg et al., 2017; Mikulewicz, 2018; Mikulewicz, 2019), suggesting
36 the need to enhance understanding of how resilience is being operationalized at the program and project
37 level and the net implications for human and ecological well-being.

38

39 This expansion of research has been accompanied by a shift in the policy context for climate action including
40 an increasingly strong link between climate actions and sustainable development. In particular, the SDGs
41 represent a near-term framework linking sustainability and human development in a manner that not only
42 addresses planetary health and human wellbeing, but also help better plan and implement mitigation and
43 adaptation actions to achieve these linked goals (Conway et al., 2015; Griscom et al., 2017; Allen et al.,
44 2018b; Roy et al., 2018; P.R. Shukla E. Calvo Buendia, 2019). The SDGs explicitly identify climate action
45 (SDG 13) among the goals needed to achieve sustainable development. Meanwhile, the text of the Paris
46 Agreement makes explicit mention of the importance of considering climate "in the context of sustainable
47 development" (Articles 2, 4, 6) or as "contributing to sustainable development" (Article 7) (Article 7,
48 UNFCCC, 2015). Similarly, sustainable development appears prominently within the text of the Sendai
49 Framework for Disaster Risk Reduction (UNDRR, 2015), and the Global Assessment Reports on Disaster
50 Risk Reduction (Undrr, 2019). At the micro-level, a growing literature recognizes that climate impacts tend
51 to exacerbate existing inequalities within societies, even at the level of gender inequalities within households
52 (Sultana, 2010; Arora-Jonsson, 2011; Carr, 2013). Thus, climate change impacts threaten even short-term
53 gains in sustainable development, which could be rolled back over longer adaptation and mitigation
54 horizons. For example, the COVID-19 pandemic is estimated to have reversed gains over the past several
55 years in terms of global poverty reduction (very high confidence) (Phillips et al., 2020; Sultana, 2021;
56 Wilhelmi et al., 2021) (Cross-Chapter Box COVID in Chapter 7), reflecting the risks posed by global,
57 systemic threats to development.

Do Not Cite, Quote or Distribute  18-13       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1

 2 The WGII AR5 Report noted that adapting to the risks associated with climate change becomes more
 3 challenging at higher levels of global warming (IPCC, 2014a). This was evidenced by contrasting impacts
 4 and adaptive capacity for 2� and 4癈 of warming. This relationship between levels of warming, climate risk,
 5 and reasons for concern (see Chapter 16) is also relevant to the concept of CRD. For example, recent
 6 literature on CRD emphasizes the urgency of climate action that achieve significant reduction in greenhouse
 7 gas emissions as well as the implementation of adaptation options that result in significant gains in human
 8 and natural system resilience (very high confidence) (Haines et al., 2017; Shindell et al., 2017; Xu and
 9 Ramanathan, 2017; Fuso Nerini et al., 2018). This was explored extensively in the IPCC's SR1.5 report in its
10 comparison of impacts associated with 1.5癈 versus 2癈 climate objectives and synergies and trade-offs
11 with the SDGs (IPCC, 2018a). However, the SR1.5 report and other literature also identified potential trade-
12 offs between aggressive mitigation and the SDGs (see also Frank et al., 2017; Hasegawa et al., 2018). This
13 indicates that while future magnitudes of warming are a fundamental consideration in climate-resilient
14 development, such development involves more than just achieving temperature targets. Rather, CRD
15 considers the possible transitions that enable those targets to be achieved including the evaluation of
16 different adaptation and mitigation options and how the implementation of these strategies interacts with
17 broader sustainable development efforts and goals. This interdependence between patterns of development,
18 climate risk, and the demand for mitigation and adaptation action is fundamental to the concept of CRD
19 (Fankhauser and McDermott, 2016). Therefore, climate change and sustainable development cannot be
20 assessed or planned in isolation of one another.

21

22 18.1.4 Assessing Climate Resilient Development

23

24 In operationalizing the aforementioned definitions of CRD and CRD pathways this chapter builds its
25 assessment around five core elements that provide insights relevant to policymakers actively pursuing the
26 integration of climate resilience into development. First, as noted above, climate change poses a potential
27 risk to the achievement of development goals, including global goals such as the SDGs, as well as
28 nationally- or locally-specific goals. Accordingly, Chapter 16's discussion of key risks, their implications for
29 the SDGs, and the options for risk management are fundamental to the pursuit of CRD. This includes the
30 opportunities for implementing adaptation, mitigation, or other risk management options. Yet, the
31 management of climate risk must be accompanied by interventions that address social and ecological
32 vulnerabilities that enhance climate risk.

33

34 Second, CRD is dependent on achieving transitions in key systems including energy, land and ecosystem,
35 urban and infrastructure, and industrial systems (very high confidence) (Box 18.1, Figure 18.3). In this
36 context, CRD links to the discussion of system transitions in the SR1.5 report (IPCC, 2018b; IPCC, 2018a).
37 However, in building on the SR1.5, here the assessment of CRD also recognizes the importance of
38 transitions in societal systems that drive innovation, preferences for alternative patterns of consumption and
39 development, and the power relationships among different actors that engage in CRD. In particular, the rate
40 at which actors can achieve system transitions has important implications for the pursuit of CRD. Transitions
41 that are slow to evolve or that are more incremental in nature may not be sufficient to enable CRD in
42 comparison with faster transitions that contribute to more fundamental system transformations.

43

44 Third, equity and social justice are consistently identified in the literature as being central to climate resilient
45 development (very high confidence; 18.1.1, 18.3.1.5, 18.4, 18.5). This includes designing and implementing
46 adaptation, resilience, and climate risk management options in a manner that promotes equity in the
47 allocation of the costs and benefits of those options. Similarly, the literature on CRD emphasizes equity
48 should be pursued in the implementation of options for greenhouse gas mitigation, transitions in energy
49 systems, and low-carbon development. This emphasis on equity is consistent with the SDGs which place an
50 emphasis on reducing inequality and achieving sustainable development for all.

51

52 Fourth, success in CRD and alignment of development interventions to CRD pathways (CRDPs) is
53 contingent on the presence of multiple enabling conditions (very high confidence, 18.4.2), that operate at
54 different scales ranging from those that provide capacity to implement specific adaptation options to those
55 that enable large-scale transformational change (Box 18.1). The qualities that describe sustainable
56 development processes (e.g., social justice, alternative development models, equity and solidarity as
57 described above and in Figure 18.1) lead to short-term outcomes and conditions, such as those represented

Do Not Cite, Quote or Distribute  18-14       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 by SDGs, that in an iterative fashion enable or constraint subsequent efforts toward CRD. For example,
 2 success or failure in achieving the SDGs or the Paris Agreement would shape future efforts in pursuit of
 3 CRD and the options available to different actors.

 4

 5 Fifth, CRD involves processes involving diverse actors, at different scales operating within an
 6 environmental, developmental, socio-economic, cultural, and political context, as typified in the SDG and
 7 the Paris Agreement negotiations (very high confidence) (Kamau et al., 2018) (18.4). The dependence of
 8 CRD on processes of negotiation and reconciliation among diverse actors and interests leads to the dismissal
 9 of the notion that there is a single, optimal pathway that captures the objectives, values, and development
10 contexts of all actors, even for a particular sector, country or region. Rather, preferences for different
11 pathways and specific actions in pursuit of those pathways will be subjected to intense scrutiny and debate
12 among diverse actors within various arenas of engagement (18.4), meaning the settings, places and spaces in
13 which key actors from government, civil society and the private sector interact to influence the nature and
14 course of development.

15

16 18.1.5 Chapter Roadmap

17

18 This chapter engages with understanding CRD and the pathways to achieving it by building on the concepts
19 introduced in Chapter 1 of this Working Group II report as well as the regional and sectoral context
20 presented in other chapters (18.5). Notably, this chapter takes off where Chapters 16 and 17 end: recognizing
21 the significance of the representative key risks for CRD as well as the decision-making context of different
22 actors who are implementing policies and practices to pursue different CRD pathways and manage climate
23 risk. Therefore, the chapter assesses options for pursuing CRD as well as the broader system transitions and
24 enabling conditions in support of CRD.

25

26 This chapter hosts three Cross-Chapter Boxes, which have their natural home here. The Cross-Chapter Box
27 on Gender, Justice and Transformative Pathways (Cross-Chapter Box GENDER) assesses literature
28 specifically on gender and climate change to uncover the importance of a justice focus to facilitate
29 transformative pathways, both toward CRD, as well as a means to achieving gender equity and social justice.
30 The Cross-Chapter Box on The Role of Indigenous Knowledge in Understanding and Adapting to Climate
31 Change (Cross-Chapter Box INDIG) highlights that achieving CRD requires confronting the uncertainty of a
32 climate change future. There are many perspectives about what future is desired and how to reach it.
33 Integrating multiple forms of knowledge is a strategy to build resilience and develop institutional
34 arrangements that provide temporary solutions able to satisfy competing interests (Grove, 2018). Indigenous
35 knowledge is proven to enhance resilience in multiple contexts (e.g., Chowdhooree, 2019; Inaotombi and
36 Mahanta, 2019). Meanwhile, Cross-Chapter Box FEASIB acts as an appendix to the WGII report,
37 synthesizing information on the feasibility associated with different adaptation options for reducing risk.

38

39 In assessing the opportunities and constraints associated with the pursuit of sustainable development, this
40 chapter proceeds in Section 18.2 to assess the links between sustainable development and climate action,
41 including examination of current patterns of development and consideration for synergies and trade-offs
42 among different strategies and options. Then, in Section 18.3, the chapter assesses five systems transitions to
43 identify the shifts in development that would enable CRD. Section 18.4 assesses the role of different actors
44 in the pursuit of CRD as well as the public and private arenas in which they engage. Section 18.5 synthesizes
45 CRD assessments from different WGII sectoral and regional chapters to identify commonalities and
46 differences. The chapter concludes in Section 18.6 with a summary of key opportunities for enhancing the
47 knowledge needed to enable different actors to pursue CRD.

48

49

50 [START BOX 18.1 HERE]

51

52 Box 18.1: Transformations in Support of Climate Resilient Development Pathways

53

54 Transformational changes in the pursuit of CRDPs involve interactions between individual, collective, and
55 systems change (see Figures 18.1�18.3). There are complex interconnections between transformation and
56 transition (Feola, 2015; H鰈scher et al., 2018), and they are sometimes used as synonyms in the literature
57 (H鰈scher et al., 2018). Much of the transitions literature focuses on how societal change occurs within

Do Not Cite, Quote or Distribute  18-15       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 existing political and economic systems. Transformations are often considered to involve deeper and more
 2 fundamental changes than transitions, including changes to underlying values, worldviews, ideologies,
 3 structures, and power relationships (G鰌el, 2016; O'Brien, 2016; Kuenkel, 2019; Waddock, 2019). Systems
 4 transitions alone are insufficient to achieve the rapid, fundamental and comprehensive changes required for
 5 humanity and planetary health in the face of climate change (high confidence). Transformative action is
 6 increasingly urgent across all sectors, systems and scales to avert dangerous climate change and meet the
 7 SDGs (Pelling et al., 2015; IPCC, 2018a; IPCC, 2021b; Shi and Moser, 2021; Vogel and O'Brien, 2021)
 8 (high confidence). The SR1.5 identified transformative change as necessary to achieve transitions within
 9 land, water and ecosystems systems; urban and infrastructural systems; energy systems; and industrial
10 systems. This box summarises key points in the transformations literature relevant to climate resilient
11 development.

12

13 Transformative actions aimed at `deliberately and fundamentally changing systems to achieve more just and
14 equitable outcomes', (Shi and Moser, 2021: 2) shift pathways towards CRD (high confidence).
15 Transformative action in the context of CRD specifically concerns leveraging change in the five dimensions
16 of development (people, prosperity, partnership, peace, planet) that drive societal choices and climate actions
17 towards sustainability (18.2.2; Figure 18.1). Climate actions that support CRD are embedded in these
18 dimensions of development; for example, social cohesion and equity, individual and collective agency, and
19 democratising knowledge processes have been identified as steps to transform practices and governance
20 systems for increased resilience (Ziervogel et al., 2016b; Nightingale et al., 2020; Colloff et al., 2021; Vogel
21 and O'Brien, 2021) (high confidence). Transformative actions toward sustainability and increased well-
22 being, which are dominant components of climate resilient development, include those that explicitly redress
23 social drivers of vulnerability, shift dominant worldviews, decolonialise knowledge systems, activate human
24 agency, contest political arrangements, and insert a plurality of knowledges and ways of knowing (G鰎g et
25 al., 2017; Fazey et al., 2018a; Brand et al., 2020; Gram-Hanssen et al., 2021; Shi and Moser, 2021). They
26 alter the governance and political economic arrangements through which unsustainable and unjust
27 development logics and knowledges are implemented (Patterson et al., 2017; Shi and Moser, 2021) by
28 shifting the goals of a system or altering the mindset or paradigm from which a system arises, e.g from
29 individualism and nature-society disconnect to solidarity and nature-society connectedness along the CRD
30 dimensions in figure 18.1, and connecting inner and external dimensions of sustainability, (G鰌el, 2016;
31 Abson et al., 2017; Wamsler and Brink, 2018; Fischer and Riechers, 2019; Horcea-Milcu et al., 2019;
32 Wamsler, 2019).

33

34 There is no blueprint for how transformation is generated. An expanding literature suggests that
35 transformation takes place through diverse modalities and context-dependent actions (O'Brien, 2021).
36 Transformation may require actions that disrupt moral or social boundaries and structures that are
37 perpetuating unsustainable systems and pathways (Vogel and O'Brien, 2021) (high confidence). Extreme
38 events and long-term climatic changes can trigger a realigning of practices, politics and knowledges (Carr,
39 2019; Schipper et al., 2020b) (high confidence). While some see opportunities for generating social and
40 political conditions needed for CRD in such actions and events (Beck, 2015; Han, 2015; Shim, 2015;
41 Mythen and Walklate, 2016; Domingo, 2018), this is not guaranteed. Climate shocks, when managed within
42 socio-political systems in ways that safeguard rather than alter practices and structures, can also reinforce
43 rather than shift the status quo (Mosberg et al., 2017; Carr, 2019; Marmot and Allen, 2020; Arifeen and
44 Nyborg, 2021) (high confidence). Further, in the absence of equitable and inclusive decision-making and
45 planning, realignments resulting from disruptive actions and events can limit inclusiveness and lead to poor
46 or coercive decision-making processes that undermine the equity and justice foundations of sustainable
47 development (Orlove et al., 2020; Shi and Moser, 2021) and lead to adverse socio-environmental outcomes
48 that generate transformations away from CRD (Vogel and O'Brien, 2021) (high confidence, see also CCP2).

49

50 Evidence for transformative actions largely exists at the community or city level. While identifying how to
51 rapidly and equitably generate transformations at a global scale has remained elusive, there is high
52 agreement but limited evidence from studies of ecosystem services that suggest facilitating a wide range of
53 locally-appropriate management decisions and actions can bring about positive global-scale outcomes
54 (Millennium Ecosystem, 2005). Diverse local efforts to transform towards sustainability in the face of
55 climate change have been observed, such as community mobilization for equitable and just adaptation
56 actions and alternative visions of societal well-being (Shi, 2020b) and farmer-led shifts in agricultural
57 production systems (Rosenberg, 2021). There has been an increase in transformative actions taking place

Do Not Cite, Quote or Distribute  18-16       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 through city-level resilience building aimed at shifting inequitable relations and opening up space for a
 2 plurality of actors (Rosenzweig and Solecki, 2018; Ziervogel et al., 2021) (high confidence).

 3

 4 Prospects for transformation towards climate resilient development increase when key governance actors
 5 work together in inclusive and constructive ways through engagement in political, knowledge-technology,
 6 ecological, economic, and socio-cultural arenas (high confidence,18.4.3). Yet, the interactions between key
 7 governance actors involve struggles and negotiations in addition to collaborations (Kakenmaster, 2019;
 8 Muok et al., 2021). Transformative actions meet resistance by precisely the political, social, knowledge and
 9 technical systems and structures they are attempting to transform (Blythe et al., 2018; Shi and Moser, 2021)
10 (high confidence). There is expanding evidence that many adaptation efforts have failed to be transformative,
11 but instead entrenched inequities, exacerbated power imbalances and reinforced vulnerability among
12 marginalized groups, and that, instead, marginalized groups and future trends in vulnerability need to be
13 placed at the center of adaptation planning (Atteridge and Remling, 2018; Mikulewicz, 2019; Owen, 2020;
14 Eriksen et al., 2021a; Eriksen et al., 2021b; Garschagen et al., 2021) (high confidence). Beyond the enablers,
15 drivers, or modalities, another question tackled in the literature is how to evaluate transformation by
16 establishing criteria for transformation assessments (Ofir, 2021; Patton, 2021; Williams et al., 2021),
17 experience-based lessons on managing transformative adaptation processes (Vermeulen et al., 2018), climate
18 policy integration (Plank et al., 2021), investment criteria (Kasdan et al., 2021), political economy analysis
19 frameworks for climate governance (Price, 2021).

20

21 [END BOX 18.1 HERE]

22

23

24 [START BOX 18.2 HERE]

25

26 Box 18.2: Visions of Climate Resilient Development in Kenya

27

28 The Government of Kenya's (GoK) ambition is to transform Kenya into a `newly industrializing, middle-
29 income country providing a high-quality life to all its citizens by 2030 in a clean and secure environment'
30 (Government of Kenya, 2008). Dryland regions in Kenya occupy 80-90 per cent of the land mass, are home
31 to 36% of the population (Government of Kenya, 2012) and contribute about 10 per cent of Kenya's Gross
32 Domestic Product (GDP) (Government of Kenya, 2012) which includes half of its agricultural GDP
33 (Kabubo-Mariara, 2009). In dryland regions, pastoralism has long been the predominant form of livelihood
34 and subsistence (Catley et al., 2013; Nyariki and Amwata, 2019). The GoK seeks to improve connectivity
35 and communication infrastructure within the drylands to better exploit and develop livestock, agriculture,
36 tourism, energy, and extractive sectors (Government of Kenya, 2018). It argues that the transformation of
37 dryland regions is crucial to enhance the development outcomes for the more than 15 million people who
38 inhabit these areas (Government of Kenya, 2016: 17) and to help the country to realize its wider national
39 ambitions including a 10 percent year on year growth in GDP (Government of Kenya, 2012). A key element
40 within this vision is the promotion and implementation of the Lamu Port South Sudan Ethiopia (LAPSSET)
41 project, a 2,000km long, 100 km wide economic and development corridor extending from Mombasa to
42 Sudan and Ethiopia (Enns, 2018). Supporters of the LAPSSET project argue that it will help achieve
43 priorities laid out in the Vision 2030 by opening up poorly connected regions, enabling the development of
44 pertinent economic sectors such as agriculture, livestock and energy, and supporting the attainment of a
45 range of social goals made possible as the economy grows (Stein and Kalina, 2019).

46

47 However, the development narrative surrounding LAPSSET remains controversial in its assumptions, not
48 least because it is being promoted in the context of a highly complex and dynamic social, economic and
49 biophysical setting (Cervigni and Morris, 2016; Atsiaya et al., 2019; Chome, 2020; Lesutis, 2020). Some of
50 the key trends driving contemporary and likely future change in dryland regions are changing household
51 organization, evolving customary rules and institutions at local and community levels, and shifting cultures
52 and aspirations (Catley et al., 2013; Washington-Ottombre and Pijanowski, 2013; Tari and Pattison, 2014;
53 Cormack, 2016; Rao, 2019). Dryland regions are also witnessing demographic growth and change in land-
54 use patterns linked to shifts in the composition of livestock (for example from grazers to browsers), a
55 decrease in nomadic and increase in semi-nomadic pastoralism, and transition to more urban and sedentary
56 livelihoods (Mganga et al., 2015; Cervigni et al., 2016; Greiner, 2016; Watson et al., 2016). At a landscape
57 level, land is becoming more fragmented and enclosed, often associated with increases in subsistence and

Do Not Cite, Quote or Distribute  18-17       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 commercial agriculture, and the establishment of conservancies and other group or private land holdings
 2 (Reid et al., 2014; Carabine et al., 2015; Nyberg et al., 2015; Greiner, 2016; Mosley and Watson, 2016). In
 3 addition, there are political dynamics associated with Kenya Vision 2030 and decentralization, the influence
 4 of international capital, foreign investors and incorporation into global markets (Cormack, 2016; Kochore,
 5 2016; Mosley and Watson, 2016; Enns and Bersaglio, 2020), as well as increasing militarization and conflict
 6 in the drylands (Lind, 2018). Allied to these social and political dynamics are ongoing processes of habitat
 7 modification and degradation and biophysical changes linked in part to climate variability (Galvin, 2009;
 8 Mganga et al., 2015). The interconnected nature of these drivers will intersect with LAPSSET in myriad
 9 ways. For example, the implementation of LAPSSET may accentuate some trends, such as increases in land
10 enclosure and a shift towards more urban and sedentary livelihoods (Lesutis, 2020). Conversely, the
11 perceived threat LAPSSET could pose to pastoral lifestyles may lead to greater visibility, solidarity and
12 strength of pastoralist institutions (Cormack, 2016).

13

14 There is a recognized need to adapt and chose development pathways that are resilient to climate change
15 whilst addressing key developmental challenges within dryland regions, notably, poverty, water and food
16 insecurity, and a highly dispersed population with poor access to services (Government of Kenya, 2012;
17 Bizikova et al., 2015; Herrero et al., 2016). The current vision for development of dryland regions comes
18 with both opportunities and threats to achieve a more climate resilient future. For example, the growth in and
19 exploitation of renewable energy resources, made possible through increased connectivity, brings climate
20 mitigation gains but also risks. These risks include the uneven distribution of costs in terms of where the
21 industry is sited compared with where benefits primarily accrue, and may exacerbate issues around water
22 and food insecurity as strategic areas of land become harder to access (Opiyo et al., 2016; Cormack and
23 Kurewa, 2018; Enns, 2018; Lind, 2018). Whilst LAPSSET will bring greater freedom of movement for
24 commodities, benefitting investors, improving access to markets and urban centers, supporting trade, or ease
25 of movement for tourists supporting economic goals, it can also result in the relocation of people and impede
26 access to certain locations for the resident populations. Mobility is a key adaptation behavior employed in
27 the short and long term to address issues linked with climatic variability (Opiyo et al., 2014; Muricho et al.,
28 2019). With modelled changes in the climate suggesting decreases in income associated with agricultural
29 staples and livestock-dependent livelihoods, development that constrains mobility of local populations could
30 retard resilience gains (Ochieng et al., 2017; ASSAR, 2018; Enns, 2018; Nkemelang et al., 2018). The likely
31 increase in urban populations and the growth in tourism and agriculture may lead to increases in water
32 demand at a time when water availability could become more constrained owing to the reliance on surface
33 water sources and the modelled increases in evapotranspiration due to rising mean temperature, more
34 heatwave days and greater percentage of precipitation falling as storms (ASSAR, 2018; Nkemelang et al.,
35 2018; USAID, 2018). These pressures could make it harder to meet basic health and sanitation goals for rural
36 and poorer urban populations, issues compounded further by likely increases in child malnutrition and
37 diarrheal deaths linked to climate change (WHO, 2016; ASSAR, 2018; Hirpa et al., 2018; Nkemelang et al.,
38 2018; Lesutis, 2020). Development must pay adequate attention to these interconnections to ensure that costs
39 and benefits of achieving climate mitigation and adaptation goals are distributed fairly within a population.

40

41 [END BOX 18.2. HERE]

42

43

44 18.2 Linking Development and Climate Action

45

46 The AR5 examined the relationship between climate and sustainable development in Chapter 13 (Olsson et
47 al., 2014) and Chapter 20 (Denton et al., 2014) in Working Group II and Chapter 4 (Fleurbaey et al., 2014)
48 in Working Group III. It concluded that dangerous levels of climate change would limit efforts to reduce
49 poverty (Denton et al., 2014; Fleurbaey et al., 2014). Since the AR5, the adoption of the Paris Agreement
50 and Agenda 2030 have demonstrated increased international consensus regarding the need to pursue climate
51 change as a component of sustainable development. For example, climate change impacts "undermine the
52 ability of all countries to achieve sustainable development" (United Nations, 2015) and can reverse or erase
53 improvements in living conditions and decades of development (Hallegatte and Rozenberg, 2017). However,
54 recent analysis shows that actions to meet the goals of the Paris Agreement can undermine progress toward
55 some SDGs (high agreement, medium evidence) (Pearce et al., 2018b; Liu et al., 2019; Hegre et al., 2020)
56 (18.2.5.3). Meanwhile efforts to achieve the SDGs can contribute to worsening climate change (high
57 agreement, medium evidence) (Fuso Nerini et al., 2018). These findings in the literature highlight the

Do Not Cite, Quote or Distribute  18-18       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 importance of identifying clear goals and priorities for both climate action and sustainable development as
 2 well as mechanisms for capitalizing on potential synergies between them and for managing trade-offs. In
 3 assessing literature relevant to the intersection between climate action and development, we first explore the
 4 implications of different patterns of development and development trajectories followed by more focused
 5 assessment of the links between development and climate risk.

 6

 7 18.2.1 Implications of Current Development Trends

 8

 9 Understanding the interactions between climate change, climate action, and sustainable development
10 necessitates consideration for the current development context in which different communities, nations, and
11 regions find themselves. For example, wealthy economies of the global North will encounter different
12 opportunities and challenges vis-�-vis climate change and sustainable development than developing
13 economies of the global South. Moreover, all economies are already following an existing development
14 trajectory that has implications for the type and scale of interventions associated with pursuing CRD and
15 managing climate risk. Some nations may experience particular challenges with reducing greenhouse gas
16 emissions due to the carbon-intensive nature of their energy systems (very high confidence) (18.3.1.1).
17 Others may experience acute challenges with adaptation due to existing vulnerability associated with poverty
18 and social inequality (very high confidence) (18.2.5.1). Overcoming such challenges is fundamental to the
19 pursuit of CRD.

20

21 While demonstrable progress has been made toward the SDGs and improving human well-being, globally
22 and in specific nations, some observed patterns of development are inconsistent with sustainable
23 development and the principles of CRD (very high confidence) (van Dooren et al., 2018; Eisenmenger et al.,
24 2020; Leal Filho et al., 2020). A significant literature, for example, links development to the loss of
25 biodiversity and the extinction crisis (Ceballos et al., 2017; Gon鏰lves-Souza et al., 2020; Oke et al., 2021).
26 Meanwhile, in human systems, indicators such as the limited convergence in income, life expectancy, and
27 other measures of well-being between poor and wealthy countries (with notable outliers such as China)
28 (Bangura, 2019), and the increase in income inequality and the decline in life expectancy and well-being in
29 rich countries (Rougoor and van Marrewijk, 2015; Alvaredo et al., 2017; Goda et al., 2017; Harper et al.,
30 2017; Goldman et al., 2018), suggest limitations of the current development paradigm to successfully deliver
31 universal human and ecological well-being, by the 2030s or even mid-century (TWI, 2019).

32

33 18.2.2 Understanding Development in Climate Resilient Development

34

35 Development in this report is defined as efforts, both formal and informal, to improve standards of human
36 well-being, particularly in places historically disadvantaged by colonialism and other features of early global
37 integration. Development is not limited to the SDGs, however these represent an internationally agreed sub-
38 set of goals. Prior IPCC reports employed development as a typological framing of the current state of a
39 given country or population (IPCC, 2014a) (Section 1.1.4). Such framings frequently rest upon measures of
40 economic activity, using them as proxies for the wider well-being of the population whose activity is
41 measured. For example, the level of gross domestic product (GDP) is often equated with levels of social
42 welfare, even though as a measure of market output it can be an inadequate metric for gauging well-being
43 over time particularly in its environmental and social dimensions (Van den Bergh, 2007; Stiglitz et al.,
44 2009).

45

46 The result of this broad framing linking economic growth to human well-being has been decades of policies,
47 programs, and projects aimed at growing economies at scales from the household to regional and global.
48 However, linking development to past and current modes of economic growth creates significant challenges
49 for CRD, as it implies that the very processes that have contributed to current climate challenges, including
50 economic growth and the resource use and energy regimes it relies upon, are also the pathways to
51 improvements in human well-being. This places climate resilience and development in opposition to one
52 another.

53

54 While there are many possible successful pathways to future development in the context of climate change,
55 history shows that pathways positive for the vast majority of people, typically induce significant impacts and
56 costs, especially on marginal and vulnerable people (Hickel, 2017). Frequently, considerations for social
57 difference and equity are side-lined in these processes, for example through the assumption that a growing

Do Not Cite, Quote or Distribute  18-19       Total pages: 197
    FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 economy lifts opportunity for all, further marginalizing those who are the most vulnerable to climate change
 2 (Matin et al., 2018; Diffenbaugh and Burke, 2019).

 3

 4 The Agenda 2030 and its 17 SDGs and 169 targets seeks to `leave no one behind' through five pillars (5Ps):
 5 People, Planet, Prosperity, Peace and Partnership (United Nations, 2015). The five pillars align with the
 6 dimensions of development that influence motion toward or away from CRD. The focus on people refers to
 7 inclusion rather than exclusion, and the extent to which people are empowered or disempowered to make
 8 decisions about their well-being, determine their futures and be in a position to assert their rights. This means
 9 being able to make decisions that determine whether people are on a pathway toward or away from CRD
10 (Figures 18.1�18.3. The focus on planet refers to protecting the planet, ensuring a balance of ecosystems,
11 biodiversity and human activities, and giving equal space and respect for its integrity. The focus on
12 prosperity refers to equity in well-being grounded in unanimity over shared goals and resources, rather than
13 individualism, and economic, social and technological progress grounded in stewardship and care, rather
14 than exploitation. The focus on partnership refers to mutual respect embedded in solidarity that recognizes
15 multiple worldviews and their respective knowledges, rather than singular or hierarchy of knowledge, and
16 acknowledges inherent nature-society connections, rather than posing nature as opposites or competitors.
17 The focus on peace emphasizes the need for just and equitable societies. These five pillars are interrelated
18 but local and national contexts situate current status differently around the world. Successful achievement of
19 Agenda 2030 is aligned with a safe climate with adequate mitigation and adaptation, and effective and
20 inclusive systems transitions. With these conditions, a high CRD world can be attained, noting that when
21 approached individually, the transformative potential of the SDGs is limited (Veland et al., 2021).

22 The need for transformational changes across sectors and scales to address the urgency and scope of action

23 needed to enable a climate resilient future in which goals like the SDGs might be realized requires attention

24 to the specific ways in which development action is defined and enacted (Box 18.1).

25

26 18.2.2.1 Development Perspectives

27

28 Development is about `improvement'. However there have been different and oftentimes conflicting

29 viewpoints on the improvement of `what' and `how' to improve. The diversity of positions has resulted in a

30 multitude of metrics to track development, some more influential than others on policy. Alternative measures

31 of development, while numerous, generally seek to nuance the connection between economic growth and

32 human well-being. Because they maintain core notions of progress and, in some cases, economic growth

33 seen in more mainstream models of development, they are less vehicles for transformation than

34 continuations of thinking and action fundamentally at odds with the needs of climate resilient development.

35 These include the Measure of Economic Welfare (Nordhaus and Tobin, 1973), the Index of Sustainable

36 Economic Welfare (Cobb and Daly, 1989), the Genuine Progress Indicator (Escobar, 1995), the Adjusted

37 Net Saving Index or the Genuine Savings Index (GSI), The Human Development Index (HDI), the

38 Inequality-adjusted Human Development Index (UNDP, 2016a), the Gender Development Index, the Gender

39 Inequality Index, and the Multidimensional Poverty Index, the Index of Sustainable Economic Welfare

40 (ISEW) (Daly and Cobb, 1989), the Genuine Progress Indicator (GPI) (Kubiszewski et al., 2013), Gross

41 National Happiness (GNH) (Ura and Galay, 2004), Measures of Australia's Progress (MAP) (Trewin and

42 Hall, 2004), the OECD Better Life Index (OECD, 2019a), and the Happy Planet Index (NEF, 2016).

43

44 In terms of their historical trajectory, different perspectives on development can be broadly divided into five

45 categories.

46  a) Development as economic growth (1950s onwards): Equating development with economic growth

47  was a natural outcome of the dominance of economics as the major discipline to study problems of

48  newly independent countries in the 1950s (Escobar, 1995), measured through GDP. Environment was

49  not a policy concern in the immediate period after decolonization. The GDP measure has withstood

50  the test of time, in spite of being an inexact measure of human well-being, and is the widely used

51  metric globally to track development. Recent improvements to GDP have tried to account for

52  environmental factors (Gundimeda et al., 2007; United Nations, 2021).

53  b) Development as distributional improvements (1970s onwards): That economic growth does not

54  automatically result in decline in poverty and improved distribution of income became apparent in the

55  1970s. Welfare measures were thus promoted that involved `redistribution with growth' (Chenery,

56  1974). These distributional concerns have re-emerged in the last two decades with the widening gap

57  between the richer and poorer groups of the population (Chancel and Piketty, 2019) and also the

    Do Not Cite, Quote or Distribute  18-20                                             Total pages: 197
    FINAL DRAFT                         Chapter 18  IPCC WGII Sixth Assessment Report

1   increased attention to `ecological distribution conflicts' (Martinez-Alier, 2021). The political economy

2   perspective, highlighting continued dependencies of countries in the Global South on the Global

3   North, now evolved into political ecology highlighting environmental concerns between and within

4   countries. Environment was not yet a policy priority, despite that the links between development and

5   environment were becoming clearer.

6   c) Development as participation (1980s onwards): Bottom-up responses emphasizing sustainable

7   livelihoods and local-level development emerged in the 1980s. The movement which involved

8   independent and uncoordinated efforts by grassroots activists, social movements and NGOs became

9   `mainstreamed' into development in the 1990s (Chambers, 2012). The multidimensional nature of

10  poverty was acknowledged at the global policy level (World Bank, 2000) and there was wider

11  acceptance of the role of non-economics social sciences as well as critical approaches in research on

12  development and poverty (Thomas, 2008). Participatory development involved decentralization and

13  local planning, emphasizing protection of local natural resources in addition to improving living

14  standards.

15  d) Development as expansion of human capabilities (1980s onwards): The human development and

16  capabilities approach was the first formidable response to the GDP-centric view of development (Sen,

17  2000; Deneulin and Shahani, 2009). Studies showed that improvements in income did not necessarily

18  improve human well-being in other dimensions such as health and education, or more broadly put,

19  `freedoms' (Ruggeri Laderchi et al., 2003). The capabilities idea was influential in global policy

20  making through Human Development Reports and metrics such as Human Development Index (HDI)

21  and Multidimensional Poverty Index (MPI). However, environmental sustainability was not a major

22  component in this approach until much later (Alkire and Jahan, 2018). Recent improvements to HDI

23  such as the Planetary pressures-adjusted HDI (United Nations, 2020) is a step in this direction.

24  e) Development as post-growth (2010 onwards): The late 1980s saw a big push towards taking the

25  environment to the center of the global policy agenda (World Commission on Environment and

26  Development, 1987). However, progress in addressing environmental questions has been slow. As

27  compared to Millennium Development Goals (MDGs), SDGs aim to tackle environmental concerns

28  by explicitly tracking progress on multiple indicators. Nevertheless, the approach in these policy

29  propositions sits largely within the economic growth framework itself. The climate change challenge

30  and the financial crisis of 2008 led many scholars, ecological economists and environmental social

31  scientists in particular, to argue for a post-growth world. Post-growth (Jackson, 2021), degrowth

32  (Kallis, 2018; Hickel et al., 2021) and other environmentalist scholarship takes inspiration from

33  critiques of development such as post-development (Escobar, 1995). The argument here is not for

34  better metrics but for imagining and working towards systemic change in the wake of the climate

35  crisis. The challenge however is how to account for historical differences in economic growth and

36  living standards between Global North and Global South and to protect the interests of Global South

37  in the spirit of `common but differentiated responsibilities' to climate change adaptation and

38  mitigation. As empirical studies in Global South have demonstrated (Lele et al., 2018), developing

39  countries face multiple stressors, climate change being just one among them, and there are multiple

40  normative concerns in developing country contexts, such as equity and justice, and not merely

41  resilience (very high confidence).

42 To achieve climate resilient development requires framings of development that move away from linear

43 paradigms of development as material progress by focusing on diversity and heterogeneity, wellbeing and

44 equality, not only in contemporary practices, but also pathways of change over time (Gibson-Graham, 2005;

45 Gibson-Graham, 2006). Such approaches, which are fundamentally aligned with ecological and ecosystem-

46 based environmental assessments which identified heterogeneity of approaches and actions as the most

47 effective path to a sustainable world (Millennium Ecosystem Assessment, 2005), emphasize the importance

48 of cultural, linguistic and religious diversity, not merely as alternative sources of information about the world,

49 but as different paradigms of well-being (Kallis, 2018). These include indigenous and local knowledges that

50 provide alternatives to these framings of the world (Cross-Chapter Box INDIG). This broad reframing of

51 development includes a focus on visions such as `buen vivir' (Cubillo-Guevara et al., 2014; Walsh, 2018;

52 Acosta et al., 2019), ecological Swaraj (Kothari et al., 2014; Demaria and Kothari, 2017; Shiva, 2017), and

53 Ubuntu (Dreyer, 2015; Ewuoso and Hall, 2019), among others. All are linked by relationships with nature

54 radically different from the Western mechanistic vision, presenting not only framings of development and the

55 environment that yield locally-appropriate climate resilient development pathways, but serve as examples of

56 alternative ways of living in balance with nature that might inform similar thinking in other places.

57

    Do Not Cite, Quote or Distribute    18-21       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 18.2.2.2. Complexity of Development and Climate Action
 2 Differing perspectives on development are in part determined by the multiple diverse priorities held by
 3 different actors and nations. Another reason is that development is not a linear process with a single goal,
 4 and active development planning requires simultaneously taking multiple processes and factors into account.
 5 This is well illustrated by growing attention to climate security. The AR5 delivered conflicting messages
 6 regarding climate change and security (Gleditsch and Nord錽, 2014), yet the understanding of climate-related
 7 security risks has made substantial progress in recent years (von Uexkull and Buhaug, 2021). Although
 8 there remains a considerable research gaps in certain regions (Adams et al., 2018), a large body of qualitative
 9 and quantitative studies from different disciplines provides new insight into the relationship of climate
10 change and security (Buhaug, 2015; De Juan, 2015; Brzoska and Fr鰄lich, 2016; Abrahams and Carr, 2017;
11 Sakaguchi et al., 2017; Moran et al, 2018; Scheffran, 2020). Though not the only cause (Sakaguchi et al.,
12 2017; Mach et al., 2019), climate change undermines human livelihoods and security, because it increases
13 the populations vulnerabilities, grievances, and political tensions through an array of indirect � at times non-
14 linear � pathways, thereby increasing human insecurity and the risk of violent conflict (van Baalen and
15 Mobj鰎k, 2018; Koubi, 2019; von Uexkull and Buhaug, 2021). Indeed, context, as well as timing and spatial
16 distribution matter and need to be accounted for (Abrahams, 2020).

17

18 In line with this better understanding, climate change and security have been reframed in the political space,
19 to focus more on human security. The solutions to climate-related security risks cannot be military, but are
20 linked to development and people's vulnerabilities in complex social and politically fragile settings
21 (Abrahams, 2020). This has resulted in integration of climate-related security risk into institutional and
22 national frameworks (Dellmuth et al., 2018; Scott and Ku, 2018; Aminga and Krampe, 2020), including
23 several NDCs (Jernn鋝 and Linn閞, 2019; Remling, 2021). One example is the UN Climate Security
24 Mechanism � set up in 2018 between UNDP, UNEP and UN DPPA to help the UN more systematically
25 address climate-related security risks and devise prevention and management strategies. Yet, work remains
26 in bridging these concerns with practical responses on the ground (Busby, 2021). Especially since emerging
27 research building on the maladaptation literature, shows that this practice cannot just mean adding adaptation
28 and mitigation to the mix of development strategies in a given location, as this may have unintended and
29 unanticipated effects and might even backfire completely (Dabelko et al., 2013; Magnan et al., 2020;
30 Mirumachi et al., 2020; Schipper, 2020; Swatuk et al., 2021). In extremely underdeveloped, fragile contexts
31 such as Afghanistan, the local-level side effects of climate adaptation and mitigation projects might result in
32 different development outcomes and question the potential for sustainable peace (Krampe et al., 2021).
33 Given the clearer understanding of the intertwined nature of climate change, security, and development �
34 especially in fragile and conflict affected regions � a rethinking of how to transfer this knowledge into policy
35 solutions is necessary for the formulation of climate resilient development.

36

37 18.2.3 Scenarios as a Method for Representing Future Development Trajectories

38

39 Sustainable development represents specific development processes and priorities that can affect climate
40 risk. As a result, sustainable development both shapes the context in which different actors experience
41 climate change and represents a potential opportunity, particularly by reducing climate risk by addressing
42 vulnerability, inequity, and shifting development toward more sustainable trajectories (IPCC, 2012; Denton
43 et al., 2014; IPCC, 2014b; IPCC, 2014a; IPCC, 2018a; IPCC, 2019b). As assessed in past IPCC special
44 reports and assessment reports, this same literature has also illustrated how different socioeconomic
45 conditions affect mitigation options and costs. For example, variations in future economic growth,
46 population size and composition, technology availability and cost, energy efficiency, resource availability,
47 demand for goods and services, and non-climate-related policies (e.g., air quality, trade) individually and
48 collectively have all been shown to result in different climates and contexts for mitigation and adaptation.

49

50 One common approach for exploring the implications of different development trajectories is the use of
51 scenarios of future socioeconomic conditions, such as the Shared Socioeconomic Pathways (SSPs) (O'Neill
52 et al., 2017). The SSPs represent sets of future global societal assumptions based on different societal,
53 technological, and economic assumptions that result in different development trajectories. Such scenarios
54 often correspond to a small set of scenario archetypes (Harrison et al., 2019; Sitas et al., 2019; Fergnani and
55 Song, 2020) in that they reflect core themes regarding the future of development such as sustainability versus
56 rapid growth. Scenarios with assumptions more closely aligned with sustainability agendas (e.g., SSP1-
57 Sustainability) commonly imply lower greenhouse gas emissions and projected climate change (see WGIII

Do Not Cite, Quote or Distribute  18-22       Total pages: 197
    FINAL DRAFT                                  Chapter 18             IPCC WGII Sixth Assessment Report

1 AR6 Chapter 3), lower mitigation costs for ambitious climate goals (see WGIII AR6 Chapter 3), lower

2 climate exposure due in large part to the size of society (see Chapter 16), and greater adaptive capacity (Roy

3 et al., 2018) (see also Chapter 16). In contrast, scenarios with rapid global economic and fossil energy

4 growth (e.g., SSP5-Fossil-Fueld Development) imply higher emissions and project climate change, higher

5 mitigation costs, as well as greater social and economic capacity to adapt to climate change impacts (Hunt et

6 al., 2012) (Table 18.1).

7

8 The SSPs incorporate various assumptions regarding population, GDP, and greenhouse gas emissions, for

9 example, that are relevant to development and climate resilience. In addition, the SSPs have been used to

10 explore a broad range of development outcomes for human and ecological systems (Table 18.1), including

11 multiple studies explore futures for food systems, water resources, human health, and income inequality.

12 Limited, top-down modelling studies have used the SSPs to explore issues such as societal resilience

13 (Schleussner et al., 2021) or gender equity (Andrijevic et al., 2020a). Such studies indicate that different

14 development trajectories have different implications for future development outcomes, but results vary

15 significantly among different climate (e.g., representative concentration pathways [RCPs]) and development

16 contexts, resulting in limited agreement among different SSPs (Table 18.1). Nevertheless, for some

17 outcomes, SSPs are associated with generally similar outcomes. Over the near-term (e.g., 2030), those

18 outcomes are strongly influenced by development inertia and path dependence, reducing differences among

19 SSPs. Outcomes diverge later in the century, but fewer studies explore futures beyond 2050. Collectively,

20 the scenarios reflect trade-offs associated with different development trajectories (Roy et al., 2018), with

21 some SSPs foreshadowing outcomes that are positive in some contexts, but negative in others (Table 18.1).

22 For example, pathways that lead to poverty reduction can have synergies with food security, water, gender,

23 terrestrial and ocean ecosystems that support climate risk management, but also poverty alleviation projects

24 with unintended negative consequences that increase vulnerability (e.g., Ley, 2017; Ley et al., 2020).

25

26 Table 18.1: Implications of different socioeconomic development pathways for CRD indicators. Studies presented in
27 the above table include qualitative storylines and quantitative scenarios for two or more SSPs. Arrows and color coding
28 reflect the positive or negative impacts on sustainability based on aggregation of results for the 2030-2050 time horizon
29 across the identified studies. Confidence language reflects the number of studies upon which results are based
30 (evidence) and the agreement among studies regarding the direction of change (agreement).

                                          Shared Socioeconomic Pathway

    Developmen     Releva   Sustainabili  Middl  Region  Inequalit        Fossil-  Confiden   References
     t Indicator   nt SDG         ty       e of     al        y            fueled      ce
                                           the                          Developme
                               (SSP1)     Road   Rialry   (SSP4)                   Evidence/
                                                 (SSP3)                      nt    Agreeme
                                          (SSP2                           (SSP5)
                                             )                                         nt

                                                                                              (Hasegawa

                                                                                              et al., 2015;

    Agriculture,                                                                              Palazzo et

    Food, &                                                                                   al., 2017;

    Forestry                                                                                  Riahi et al.,
    � Agriculture
                                                                                   Low 2017; Duku

    production                                                                     Agreemen et al., 2018;

    � Forestry     SDG 2    &             1      (           (          (          t/         Chen et al.,

    production                                                                     Robust 2019;

    � Food                                                                         Evidence Daigneault

    security                                                                                  et al., 2019;

    � Hunger                                                                                  Mitter et

                                                                                              al., 2020;

                                                                                              Mora et al.,

                                                                                              2020)

                                                                                              (Chen et al.,

    Health &                                                                                  2017; Mora

    Well-Being                                                                     Medium et al., 2017;

    � Exces                           1   1      1           (          (          Agreemen Aleluia
     s mortality SDG 3                                                             t/ Robust Reis et al.,

    � Air                                                                          Evidence 2018;

    quality                                                                                   Asefi-

                                                                                              Najafabady

    Do Not Cite, Quote or Distribute             18-23                                 Total pages: 197
FINAL DRAFT                          Chapter 18     IPCC WGII Sixth Assessment Report

  � Vecto                                                                                       et al., 2018;
   r-borne
   disease                                                                                      Chen et al.,
  � Life
   Satisfaction                                                                                 2018;

                                                                                                Harrington

                                                                                                and Otto,

                                                                                                2018;

                                                                                                Marsha et

                                                                                                al., 2018;

                                                                                                Sellers and

                                                                                                Ebi, 2018;

                                                                                                Ikeda and

                                                                                                Managi,

                                                                                                2019;

                                                                                                Rohat et al.,

                                                                                                2019;

                                                                                                Wang et al.,

                                                                                                2019; Chae

                                                                                                et al., 2020)

                                                                                                (Wada et

                                                                                                al., 2016)

Water &                                                                                         (van

Sanitation                                                                                      Puijenbroek
� Water use
                                                                                     High et al., 2014;

� Sanitation SDG 6       &        (  (           1  1                                Agreemen Yao et al.,

access                                                                               t/ Medium 2017)

� Sewage                                                                             Evidence (Mouratiad

discharge                                                                                       ou et al.,

                                                                                                2016;

                                                                                                Graham et

                                                                                                al., 2018)

                                                                                                (Rao et al.,

                                                                                                2019b;

                                                                                     Medium Emmerling

Inequality                                                                           Agreemen and Tavoni,
                                                                                     t/ Limited 2021;
� Gini SDG 10            &        &  &           1  &

coefficient                                                                          Evidence Gazzotti et

                                                                                                al., 2021)

Ecosystems                                                                                      (Li et al.,

and                                                                                             2017; Chen

Ecosystem                                                                                       et al., 2019;

Services                                                                                        Li et al.,
� Aquatic
                                                                                                2019b;

resources                                                                            High Chen et al.,

� Urban                                                                                             2020b;
                                                                                     Agreemen
expansion SDG 14         (        (  (           (  (
                                                                                                    Song et al.,
                                                                                     t/ Medium
� Habitat        SDG 15                                                              Evidence 2020b;

provision                                                                                       McManama

� Carbon                                                                                        y et al.,

sequestrati                                                                                     2021;

on                                                                                              Pinnegar et

� Biodiversit                                                                                   al., 2021)

y

Legend
$ Balance of studies suggest large increasing threat to sustainable development
( Balance of studies suggest moderate increasing threat to sustainable development
1 Studies suggest both threats and benefits to sustainable development
& Balance of studies suggest moderate increasing benefit to sustainable development
# Balance of studies suggest large increasing benefit to sustainable development

Table Notes:

Studies presented in the above table include qualitative storylines and quantitative scenarios for two or more SSPs.

Arrows and color coding reflect the positive or negative impacts on sustainability based on aggregation of results for

Do Not Cite, Quote or Distribute     18-24                                           Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

         the 2030-2050 time horizon across the identified studies. Confidence language reflects the number of studies upon

         which results are based (evidence) and the agreement among studies regarding the direction of change (agreement).

 1

 2 While the scenarios literature is useful for characterizing the potential climate risk implications of different
 3 global societal futures, important limitations impact their use in climate risk management planning (very
 4 high confidence). The first is the often highly geographically aggregated nature of the SSPs and other
 5 scenarios, which, in the absence of application of nesting or downscaling methods, often lack regional,
 6 national, or sub-national context, particularly regarding social and cultural determinants of vulnerability (van
 7 Ruijven et al., 2014). Furthermore, there is limited understanding of the cost and what is required to
 8 transform from today into each socioeconomic future, or the opportunity to shift from one pathway to
 9 another (18.3). Furthermore, the characteristics of the pathways suggest that they are not equally likely, there
10 are relationships implied in assumptions that are uncertainties to consider (e.g., land productivity
11 improvements are land saving), it is difficult to identify the role of different development characteristics, and
12 policy implementation is stylized. In general, global assessments are not designed to inform local planning
13 given that there are many local circumstances consistent with a global future and unique local development
14 context and uncertainties to manage--demographic, economic, technological, cultural, policy.

15

16 Overall, pursuing sustainable development in the future is shown to have synergies and trade-offs in its
17 relationships with every element of climate risk: the emissions and mitigation determining hazard, the size,
18 location, and composition of development determining exposure; and the adaptive capacity determining
19 vulnerability. Importantly, the scenarios literature overall has found trade-offs such that none of the global
20 societal projections achieve all the sustainable development goals (very high confidence) (Roy et al., 2018)
21 (18.2.5.3). Historical evidence supports this as well, for example, finding low-cost energy and food access
22 historically associated with higher emissions but greater adaptive capacity, and energy efficiency innovation
23 contributing to lower emissions and greater adaptive capacity (e.g., Blanford et al., 2012; Blanco et al., 2014;
24 Mbow et al., 2019; USEPA, 2019). The literature suggests that trade-offs in the pursuit of sustainable
25 development are inevitable. Managing those trade-offs, as well as capitalizing on the synergies, will be
26 important for CRD, particularly given trade-offs have distributional implications that could contribute to
27 inequities (18.2.5.3).

28

29 18.2.4 Climate Change Risks to Development

30

31 Over the next decade, additional climate change is expected regardless of the scale of greenhouse gas
32 mitigation efforts (IPCC, 2021a). Across the global scenarios analyzed in the AR6, global average
33 temperature changes relative to the reference period 1850-1900 range from 1.2癈 to 1.9癈 for the period
34 2021�2040 and 1.2癈 to 3.0癈 for the period 2041-2060 (WGI AR6 SPM very likely range). However, the
35 feasibility of emissions pathways (particularly, RCP8.5) affect the plausibility of the associated climate
36 projections, potentially lowering the upper end of these ranges (see WGIII AR6 Chapter 3). There is
37 significant overlap between climate scenario ensemble ranges from different emissions scenarios through
38 2050, more so than through 2100 (Lee et al., 2021). There is also overlap between emissions scenario
39 ensembles consistent with different temperature outcomes (see WGIII AR6 Chapter 3). Emissions pathway
40 ranges represent uncertainties for policy-makers and organizations to consider and manage (Rose and Scott,
41 2018, 2020) regarding, among other things, economic growth and structure, available technologies, markets,
42 behavioral dynamics, policies, and non-CO2 climate forcings (see WGIII AR6 Chapter 3), while climate
43 pathway ranges represent bio-physical climate system and carbon cycle uncertainties (Lee et al., 2021). For
44 all climate projections and variables, there is significant regional heterogeneity and uncertainty in projected
45 climate change (very high confidence) (IPCC, 2021a). Figure 18.4 (left panel) presents examples for average
46 and extreme temperature precipitation change (see also 18.5 and Tables 18.4�18.5 for more regional detail).
47 Similarly, for all emissions projections, there is significant regional, sectoral, and local heterogeneity and
48 uncertainty regarding potential pathways for climate action (see WGIII AR6 Chapter 3 and Chapter 4). Not
49 all uncertainties are represented in projected emissions pathway ensembles, such as policy timing and design
50 (e.g., Rose and Scott, 2018) or climate projection ensembles.

51

52 The projected ranges for near-term and mid-term global average warming levels are estimated to result in
53 increasing key risks and reasons for concern (Chapter 16). Chapter 16 developed aggregate "Representative
54 Key Risks" (RKRs) as indicators for subsets of approximately one hundred sectoral and regional key risks
55 indicators. The RKRs include risks to coastal socio-ecological systems, terrestrial and ocean ecosystems,

Do Not Cite, Quote or Distribute  18-25       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 critical physical infrastructure, networks and services, living standards and equity, human health, food
 2 security, water security, and peace and migration. The majority of these risks are directly linked to
 3 sustainable development priorities and the SDGs (Chapter 16, WGII AR6 sectoral and regional chapters;
 4 (Roy et al., 2018; IPCC, 2019d; IPCC, 2019b). Therefore, climate risks represent a potential additional
 5 challenge to pursuing sustainable development priorities, but also potential opportunities due to geographic
 6 variation in climate impacts. In addition, positive synergies have been found between sustainable
 7 development and adaptation, but trade-offs are also possible (e.g., Roy et al., 2018).

 8

 9 For all RKRs, additional global average warming is expected to increase risk. However, the increases vary
10 significantly by RKR, and across the underlying key risks represented within each RKR. Geographic
11 variation in key risk implications is only partially assessed in Chapter 16, but evidence can be drawn from
12 the WGII individual regional chapters. Regionally, key risks are found to be potentially greatest in
13 developing and transition economies (Chapter 16 and sectoral chapters), which is also where the least-cost
14 emissions reductions are shown to be (see WGIII AR6 Chapter 3). See Figure 18.4 for an example of key
15 risk geographic heterogeneity (see also 18.5 for regional detail). Chapter 16 also maps the RKRs to an
16 updated aggregate "Reasons for Concern" (RFC) framing. Thus, increasing RKR risk implies increasing
17 RFC associated with unique and threatened systems, extreme weather events, distribution of impacts, global
18 aggregate impacts, and large-scale singular events.

19

20 Climate risks are found to vary with future warming levels, the development context and trajectory, as well
21 as by the level of investment in adaptation. Together, these three dimensions define risk � with projected
22 climate changes defining the hazard, development defining the exposure, and development and adaptation
23 defining vulnerability. However, how these different dimensions interact and the level of scientific
24 understanding vary significantly among different types of risk. For human systems, in general, the poor and
25 marginalized are found to have greater vulnerability for a given hazard and exposure level. With some level
26 of global average warming expected regardless of mitigation efforts, human and natural systems will be
27 exposed to new conditions, but some level of adaptation should also be expected.

28

29

30

31 Figure 18.4: Regional projected select climate change and sustainable-development-related climate impact variables by
32 global warming level. Sources: WGI and WGII AR6 reports.

33
34

35 18.2.5 Options for Managing Future Risks to Climate Resilient Development

36

Do Not Cite, Quote or Distribute  18-26       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 The pursuit of CRD requires not only the implementation of individual adaptation, mitigation, and
 2 sustainable development initiatives, but also their careful coordination and integration. This section assesses
 3 the literature on CRD in the context of key climate change risks (Chapter 16); gaps in adaptation that
 4 contribute to risk; potential synergies and trade-offs among mitigation, adaptation and sustainable
 5 development; and the mechanisms for managing those trade-offs.

 6

 7 18.2.5.1 Adaptation

 8

 9 18.2.5.1.1 Adaptation and climate-resilient development
10 Given adaptation is recognized as a key element of addressing climate risk and CRD, the capacity for
11 adaptation implementation is an important consideration for CRD. The AR5 noted a significant overlap
12 between indicators of sustainable development and the determinants of adaptive capacity, and suggested that
13 adaptation presents an opportunity to reduce stresses on development processes and the socio-ecological
14 foundations upon which they depend (Denton et al., 2014). At the same time, it also noted that building
15 adaptive capacity for sustainable development might require transformational changes that shift impacted
16 systems to new patterns, dynamics, or places (Denton et al., 2014). Thus, adaptation interventions and
17 pathways can further the achievement of development goals such as food security (Campbell et al., 2016;
18 Douxchamps et al., 2016; Richardson et al., 2018; Bezner Kerr et al., 2019) and improvements in human
19 health (Watts et al., 2019) including in systems where animals and humans live in close proximity (very high
20 confidence) (Zinsstag et al., 2018). However, to do so requires not only the avoidance of incremental
21 adaptation actions that extend current unsustainable practices, but also the ability to manage and overcome
22 the barriers which arise when the limits of incremental adaptation are reached (high agreement; medium
23 evidence) (Few et al., 2017; Vermeulen et al., 2018; Fedele et al., 2019).

24

25 Since AR5, the scientific community has deepened its understanding of the relationship between adaptation
26 and sustainable development (very high confidence), particularly with regard to the place of resilience at the
27 intersection of these two arenas. The literature has moved forward in its identification of specific overlaps in
28 sustainable development indicators and determinants of adaptive capacity, how adaptation might reduce
29 stress on development processes and their socio-ecological foundation, and how building adaptive capacity
30 might facilitate needed transformative changes. Broadly speaking, work on these topics comes from one of
31 two perspectives. One perspective speaks to adaptation practices that might further sustainable development
32 outcomes, while another perspective draws on deeper understandings of the socio-ecological dynamics of the
33 systems in which we live, and which we may have to transform in the face of climate change impacts. These
34 two literatures are not yet well-integrated, leaving gaps in our knowledge of how best to implement
35 adaptation in a manner that achieves sustainable development.

36

37 The literature considering adaptation and development in practice since AR5 suggests that efforts to connect
38 adaptation to sustainable development should address proximate and systemic drivers of vulnerability (Wise
39 et al., 2016) while remaining flexible and reversable to avoid the lock-in of undesirable or mal-adaptive
40 trajectories (Cannon and M黮ler-Mahn, 2010; Wise et al., 2016). Such goals require critical reflection on
41 processes for decision-making and learning. In the AR5, more inclusive, participatory adaptation processes
42 were presumed to benefit development planning by including a wider set of actors in discussions of future
43 goals (Denton et al., 2014). The post-AR5 literature expands on these critical perspectives to provide context
44 regarding when participation is most effective. For example, (Eriksen et al., 2015) emphasize the need to
45 build participatory adaptation processes to avoid subsuming adaptation goals to development-as-usual while
46 (Kim et al., 2017b) argues that this practice is most effective when it is focused on development efforts and
47 considers how climate change will challenge the goals of those efforts. Adaptation, while presenting an
48 opportunity to foster transformations needed to address the impacts of climate change on human well-being,
49 is also a contested process that is inherently political (medium agreement, medium evidence) (Eriksen et al.,
50 2015; Mikulewicz, 2019; Nightingale B鰄ler, 2019; Eriksen et al., 2021b). How adaptation can challenge
51 development and create a situation where CRD effectively becomes transformative adaptation, adaptation
52 that generates transformation of broader aspects of development, remains unclear (medium agreement,
53 limited evidence) (Few et al., 2017; Schipper et al., 2020c).

54

55 The critical literature on socio-ecological resilience, which has grown substantially since the last AR (very
56 high confidence), speaks to some of these questions. Since AR5, the IPCC and the wider literature on socio-
57 ecological resilience have shifted their use of the term to reflect not only the capacity to cope with a

Do Not Cite, Quote or Distribute  18-27       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 hazardous event or trend or disturbance, but also the ability to adapt, learn, and transform in ways that
 2 maintains a socio-ecology's essential function, identity and structure (WGII Chapter 1, Glossary). This
 3 change in usage is significant in that it shifts resilience from an emergent property of complex socio-
 4 ecological systems to a deeply human product of efforts to manage ecology, economy, and society to
 5 specific ends. This definition of resilience recognizes the need to define what is an essential identity,
 6 function, and structure for a given system, questions rooted not in ecological dynamics, but in politics,
 7 agency, difference, and power that emerge around the management of ecological dynamics (Cote and
 8 Nightingale, 2011; Brown, 2013; Cretney, 2014; Forsyth, 2018; Matin et al., 2018; Carr, 2019).

 9

10 By connecting this framing of socio-ecological dynamics to the literature on the principles for adaptation
11 efforts that meet development goals, new work has begun to identify 1) how adaptation can reduce stress on
12 development processes, 2) how it might facilitate transformative change, and 3) where adaptation
13 interventions might either drive system rigidity and precarity, or otherwise challenge development goals
14 (Castells-Quintana et al., 2018; Carr, 2020). For example, Jordan (2019) draws upon these contemporary
15 framings of resilience to highlight the ways in which coping strategies perpetuate the gendered norms and
16 practices at the heart of women's vulnerability in Bangladesh. Forsyth (2018) draws upon this work to
17 highlight the ways in which the theory of change processes used by development organizations tend to
18 exclude local experiences and sources of risk, and thus foreclose the need for transformative pathways to
19 achieve development goals. Carr (Carr, 2019; 2020) draws upon evidence from sub-Saharan Africa to
20 develop more nuanced understandings of the ways in which different stressors and interventions either
21 facilitate or foreclose transformative pathways, while pointing to the existence of yet poorly-understood
22 thresholds for transformation in systems that can be identified and targeted by interventions.

23

24 18.2.5.1.2 Adaptation Gaps
25 Adaptation gaps are defined as "the difference between actually implemented adaptation and a societally set
26 goal, determined largely by preferences related to tolerated climate change impacts and reflecting resource
27 limitations and competing priorities" (UNEP, 2014; UNEP, 2018a). Adaptation deficit is a similar concept,
28 described as an inadequate or insufficient adaptation to current conditions (see Ch 1). Adaptation gaps or
29 deficits arise from a lack of adequate technological, financial, social, and institutional capacities to adapt
30 effectively to climate change and extreme weather events, which are in turn linked to development (very
31 high confidence) (Fankhauser and McDermott, 2014; Milman and Arsano, 2014; Chen et al., 2016; Asfaw et
32 al., 2018) (18.2.2).

33 Currently, there is no consensus around approaches to assess the effectiveness of adaptation actions across
34 contexts and therefore measure adaptation gaps at a global scale (Singh et al., 2021a). UNEP (2021) suggests
35 that comprehensiveness, inclusiveness, implementability, integration and monitoring and evaluation can be
36 used to assess them (see also Cross-Chapter Box FEASIB). However, limited information is available about
37 future trends in national-level adaptation, and the development of monitoring and evaluation mechanisms.
38 Despite the challenges of measurement associated with adaptation gaps, available evidence from smaller
39 scales across several regions, communities, and businesses suggest that significant adaptation gaps have
40 existed in historical contexts of climate change, while expectations of extreme heat, increasing storm
41 intensity, and rising sea levels will create the context for the emergence of new gaps (very high confidence)
42 (Hallegatte et al., 2018; UNEP, 2018a; Dellink et al., 2019; UNEP, 2021). These adaptation gaps create risks
43 to well-being, economic growth, equity, the health of natural systems, and other societal goals. The negative
44 impacts of these gaps can be compounded by adaptation efforts that are considered maladaptive or by
45 development actions that are labelled as adaptation (see Chapter 16).

46 A higher level of adaptation finance is critical to enhance adaptation planning and implementation and
47 reduce adaptation gaps, particularly in developing countries (very high confidence) (UNEP, 2021) (Cross-
48 Chapter Box FINANCE in Chapter 17, 18.4.2.2). However, adaptation finance is not keeping pace with the
49 rising adaptation costs in the context of increasing and accelerating climate change, as "annual adaptation
50 costs in developing countries alone are currently estimated to be in the range of US$70 billion, with the
51 expectation of reaching US$140�300 billion in 2030 and US$280�500 billion in 2050" (UNEP, 2021).
52 Investment in attaining SDGs helps bridge adaptation gaps (Birkmann et al., 2021), but care needs to be
53 taken to avoid maladaptation through mislabeling. Integration of the indigenous and local knowledge
54 systems is anticipated to reduce existing adaptation gaps and secure livelihood transitions.

Do Not Cite, Quote or Distribute  18-28       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1 Analysis of investments by four major climate and development funds (the Global Environment Facility,
2 the Green Climate Fund, the Adaptation Fund and the International Climate Initiative) by UNEP (2021)
3 suggests that support for green and hybrid adaptation solutions has been increasing over the past two
4 decades. These could be effective at reducing climate risks and bridging adaptation gaps while
5 simultaneously bringing important additional benefits for the economy, environment, livelihoods (UNEP,
6 2021) (see also Cross-Chapter Box NATURAL in Chapter 2).

 7 Lately, the evidence of adaptation activity in the health sector has been increasing (Watts et al., 2019), yet
 8 substantial adaptation gaps persist (UNEP, 2018a; UNEP, 2021), including gaps in humanitarian response to
 9 climate-related disasters (Watts et al., 2019). It is the under-investment in climate and health research in
10 general and health adaptation in particular that has led to adaptation gaps in the health sector (Ebi et al.,
11 2017).

12 Costs of implementing efficient adaptation measures and water-related infrastructure in water-deficient
13 regions have received attention at the global and regional level to bridge the `adaptation gap' (Hallegatte et
14 al., 2018; UNEP, 2018a; Dellink et al., 2019; UNEP, 2021). Livelihood sustainability the drylands, which
15 cover more than 40% of land surface area, are home to roughly 2.5 billion people, and support
16 approximately 50% of the livestock and 45% of the food production, is threatened by a complex and
17 interrelated range of social, economic, and environmental changes that present significant challenges to rural
18 communities, especially women (Abu-Rabia-Queder and Morris, 2018; Gaur and Squires, 2018). Adaptation
19 deficits in arid and semi-arid regions are of high order (see CCP 3). In order to reduce adaptation deficit in
20 arid and semi-arid regions comprehensive and efficient adaptation interventions integrating better water
21 management, use of non-traditional water sources, changes in reservoir operations, soil ecosystem
22 rejuvenation, and enhanced institutional effectiveness are needed (18.5) (Makuvaro et al., 2017; Mohammed
23 and Scholz, 2017; Morote et al., 2019). Communities facing the lack of adequate technological, financial,
24 human, and institutional capacities to adapt effectively to current and future climate change often encounter
25 adaptation deficits. In order to address current adaptation barriers and adaptation deficits, there is a need to
26 promote efficient adaptation measures, coupled with inclusive and adaptive governance involving
27 marginalized groups such as indigenous communities and women.

28 Although unevenly distributed urban adaptation gaps exist in all world regions (see Chapter 6). Such gaps
29 are higher in the urban centers of the poorer nations. Chapter 6 identified the critical capacity gaps at city
30 and community levels that are responsible for adaptation gaps are: "ability to identify social vulnerability
31 and community strengths, and to plan in integrated ways to protect communities, alongside the ability to
32 access innovative funding arrangements and manage finance and commercial insurance; and locally
33 accountable decision-making with sufficient access to science, technology and local knowledge to support
34 the application of adaptation solutions at scale".

35 Insufficient financial resources are the main reasons for the coastal adaptation gap particularly in the Global
36 South (see CCP2). Engaging the private sector with a range of financial tools is crucial to address such gaps
37 (see CCP2). An urgent and transformative action to institutionalize locally-relevant integrative adaptation
38 pathways is crucial for closing coastal adaptation gaps. Additional efforts are in place for assessing global
39 adaptation progress (see Cross-Chapter Box PROGRESS in Chapter 17].

40 18.2.5.1.3 Adaptation implementation
41 As discussed in Chapter 16, adaptation is a key mechanism for managing climate risks (Chapter 16), and
42 therefore for pursuing CRD. The lower estimates in Table 18.2 are associated with higher levels of
43 adaptation and more conducive development conditions. Furthermore, additional adaptation demand is
44 associated with greater levels of climate change. Adaptation is a broad term referring to many different
45 levels of response and options for natural and human systems, from individuals, specific locations, and
46 specific technologies, to nations, markets, global dynamics, and strategies at the system level. Adaptation
47 also includes endogenous reflexive and exogenous policy responses. Perspectives on limits to adaptation,
48 synergies, trade-offs, and feasibility therefore depend on where the boundaries are drawn and the objective.
49 Overall, there are a broad range of adaptation options relevant to reducing risks posed by climate change to
50 development. However, current understanding of how such options are implemented in practice, their
51 effectiveness across a range of possible climate futures, and their potential limits, is modest.

52

Do Not Cite, Quote or Distribute  18-29       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 Past assessments have evaluated individual adaptation options in terms of economic, technological,
 2 institutional, socio-cultural, environmental/ecological, and geophysical feasibility (de Coninck et al., 2018).
 3 This analysis has been updated for AR6 (Cross-Chapter Box FEASIB). These assessments identify types of
 4 barriers that could affect an option's feasibility. Among other things, this work finds that every adaptation
 5 option evaluated had at least one feasibility dimension that represented a barrier or obstacle. The barriers
 6 also imply that there are trade-offs in these feasibility dimensions to consider. Overall, insights from this
 7 work are high-level and difficult to apply to a specific adaptation context. The feasibility and ranking of
 8 adaptation opportunities, as well as the list of opportunities themselves, for a given location will vary from
 9 location-to-location, with different criteria and weighting of criteria that reflect the relevant social priorities
10 and differences in markets, technology options, and policies for managing risks and trade-offs. Integrated
11 evaluation of criteria and options is needed, that accounts for the relevant geographic context and
12 interactions between options and systems (18.5).

13

14 Sustainable development is regarded as generally consistent with climate change adaptation, helping build
15 adaptive capacity by addressing poverty and inequalities and improving inclusion and institutions (Roy et al.,
16 2018). Some sustainable development strategies could facilitate adaptation effectiveness by addressing wider
17 socio-economic barriers, addressing social inequalities, and promoting livelihood security (Roy et al., 2018).
18 With a common goal of reducing risks, sustainable development and adaptation are relatively synergistic.
19 However, trade-offs have been found and important to consider and potentially manage. Synergies have been
20 found between adaptation and poverty reduction, hunger reduction, clean water access, and health; while,
21 trade-offs have also been found, particularly when adaptation strategies prioritize one development objective
22 (e.g., food security or heat-stress risk reduction) or promote high-cost solutions with budget allocation and
23 equity implications (Roy et al., 2018) (18.2.5.3, 18.5). There are also opportunities for managing the trade-
24 offs, in particular distributional effects--by recognizing that there are trade-offs and considering alternatives
25 and complementary strategies to offset the trade-offs (Section 18.2.5.3).

26

27

28 [START BOX 18.3 HERE]

29

30 Box 18.3: Climate Resilient Development in Small Islands

31

32 Small Islands are particularly vulnerable to climate change and many are already pursuing climate resilient
33 development pathways that enable integrated responses (Allen et al., 2018a; Mycoo, 2018; Hay et al., 2019;
34 Robinson et al., 2021). Countries, such as Belize, have opted for a systems-approach and are working across
35 the SDGs to increase integration (Allen et al., 2018a). This includes rethinking disaster reconstruction
36 mechanisms in the Caribbean and introducing more diversified and sustainable tourism economies that can
37 better withstand external shocks such as disruptions and loss of markets from COVID-19 (Sheller, 2021). In
38 the Seychelles, various government and tourism industry initiatives are focused on the promotion of
39 sustainable tourism ventures that lower emissions, protect and promote biodiversity conservation (e.g. new
40 marine protected areas with mitigation and adaptation benefits), and are climate resilient (Robinson et al.,
41 2021). In 2016 the Seychelles signed the world's first nature-for-debt swap wherein an NGO (The Nature
42 Conservancy) agreed to pay off Seychelles' public debt to the Paris Club (foreign creditors) in return for the
43 Seychelles government establishing marine conservation areas (Silver and Campbell, 2018).

44

45 One key area where enhanced climate risk integration is critical is infrastructure-related decisions especially
46 on coastal areas (World Bank, 2017). However, despite increasing awareness of climate risks and
47 experienced impacts, decisions on for example infrastructure locations still reflect cultural preferences. For
48 example, Hay et al. (2019) report that despite recommendations to relocate the redevelopment site of the
49 Parliamentary Complex in Samoa away from the coast, multiple cultural and historical factors influenced the
50 decisions to redevelop at the original site. In the Solomon Islands, however, emerging evidence suggests that
51 adaptation efforts to enhance the resilience of infrastructure are also serving to help urban areas address
52 problems associated with rapid urbanization and provide new opportunities for sustainable development
53 (Robinson et al., 2021).

54

55 Energy system transitions in small islands can produce synergies with SDG implementation, and can lead to
56 transformational outcomes. The Pacific island territory of Tokelau has demonstrated a nationwide energy
57 transition, sourcing 100% of their energy needs from solar power (Michalena and Hills, 2018), and many

Do Not Cite, Quote or Distribute  18-30       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 other countries such as Fiji, Niue, Tuvalu, Vanuatu, Solomon Islands and Cook Islands also have 100%
 2 renewable energy targets. Benefits of small island distributed energy systems (such as solar photovoltaic
 3 (PV) systems) include less need for large, centralized infrastructure; reduced reliance on volatile fossil fuel
 4 markets; enhanced international climate negotiations power and enhanced local job markets/skills (Dornan,
 5 2015; Cole and Banks, 2017; Weir, 2018). Additionally, renewable systems can enhance resilience to hydro-
 6 meteorological disasters (Weir and Kumar, 2020). For example, well secured ground based PV systems
 7 withstood cyclones in the Pacific island of Tonga during cyclone Gita and across the Caribbean during
 8 Hurricane Maria with power restored in days rather than weeks associated with more centralized systems
 9 (Weir and Kumar, 2020). Yet, a multitude of challenges remain. In the Pacific islands region, these include:
10 the high up front capital investment of renewables; lack of private sector investment; limited renewable
11 energy data for policy making; land tenure/rent costs; ongoing infrastructure maintenance skills and
12 requirements; political turnover; failed experimentation; difficulty in obtaining and transporting replacement
13 parts and a highly corrosive environment for equipment (Dornan, 2015; Cole and Banks, 2017; Lucas et al.,
14 2017; Weir, 2018; Weir and Kumar, 2020). The example of Pacific energy transitions demonstrates that a
15 nuanced and context specific analysis of synergies and trade-offs for energy transitions is required in order to
16 lessen the impact on fragile economies and maximize benefits for remote populations.

17

18 Labor migration is increasingly recognized as a significant factor that can contribute to climate resilient
19 development pathways for small islands. In the Pacific Islands region, labor mobility schemes are already
20 allowing for climate change adaptation and economic development to occur in labor migrants' countries of
21 origin (Smith and McNamara, 2015; Klepp and Herbeck, 2016; Dun et al., 2020). Dun et al. (2020)
22 demonstrates that temporary or circular migrants from the Solomon Islands, working in Australia under its
23 Seasonal Worker Program (similar programs operate in other developed countries), are using the money they
24 earn to invest in adaptation and development activities back home. Similarly, labor migrants from Vanuatu,
25 Kiribati, and Samoa contribute to development and in-situ climate change adaptation (at a household,
26 village, and regional level) that enable discussions about more resilient futures for their countries (Barnett
27 and McMichael, 2018; Parsons et al., 2018).

28

29 [END BOX 18.3 HERE]

30

31

32 [START BOX 18.4 HERE]

33

34 Box 18.4: Adaptation and the Sustainable Development Goals

35

36 The achievement of the SDGs represents near-term positive sustainability as well as indicating the quality
37 of development processes and actions (inclusion and social justice, degrowth and alternative development
38 models, planetary health, well-being, equity, solidary, plural knowledges and human-nature connectivity)
39 that enable CRD in the long term (18.2.2.2, 18.2.5.3). A key question is the extent to which adaptation
40 actions (or non-action) may contribute to (or undermine) SDG achievement, and in particular to shift the
41 quality of development processes and engagement within the political, economic, ecological, socio-
42 ethical and knowledge-technology arenas and hence contribute to CRDPs. Here, the relationship between
43 adaptation and SDGs is illustrated through an examination of SDG3 good health and well-being and
44 SDG16 peace, justice and strong institutions. These two are foundational to social equity and justice that
45 underpin sustainability outcomes as well as enablers of CRD.

46

47 Table Box 18.4.1 (below) provides a set of examples of how adaptation actions can either contribute to or
48 undermine SDG achievement, for SDGs 2, 3, 6, 11 and 16. In general, evidence suggests positive effects
49 of formal interventions as well as household and community-based adaptation strategies on discrete social
50 variables among target populations, particularly if they are shaped by the local context and needs, with
51 real participation and leadership by target populations (Remling and Veitayaki, 2016; Buckwell et al.,
52 2020; McNamara et al., 2020; Owen, 2020). For example, integrated adaptation approaches to the Water-
53 Energy-Food (WEF) Nexus aiming to build resilience in those sectors can lead to increased resource use
54 efficiency and coherent strategies for managing the complex interactions and tradeoffs among the water,
55 energy and food SDGs (Mpandeli et al., 2018; Nhamo et al., 2020). One such approach could involve
56 cultivating indigenous crops suited to harsh growing conditions, which would allow for agricultural
57 expansion for food and energy without increased water withdrawals (Mpandeli et al., 2018). Overall,

Do Not Cite, Quote or Distribute  18-31       Total pages: 197
    FINAL DRAFT                                      Chapter 18  IPCC WGII Sixth Assessment Report

1 adaptation commitments aiming to build resilience of vulnerable populations have typically shown to

2 contribute to SDGs focused on ending extreme poverty (SDG 1), improving food security (SDG 2),

3 improving access to water (SDG 6), ensuring clean energy (SDG 7), tackling climate change (SDG 13)

4 and halting land degradation and deforestation (SDG 15) (Antwi-Agyei et al., 2018).

5

6 However, evidence also suggests limitations of adaptation actions, with the objectives and actions often

7 being too narrow to address social justice and enable CRD. As such, adaptation actions can sometimes

8 undermine SDG achievement through exacerbating social vulnerability, inequity and uneven power

9 relations (Antwi-Agyei et al., 2018; Atteridge and Remling, 2018; Paprocki, 2018; Mikulewicz, 2019;

10 Satyal et al., 2020; Scoville-Simonds et al., 2020). This is due to adaptation practices often not accounting

11 for the differentiated ways in which minority groups are especially vulnerable. For example, designs of

12 emergency shelters should consider the fear of social stigma or abuse faced by women and girls (Pelling

13 and Garschagen, 2019).

14

15 Such maladaptive adaptation practices can undermine SDG achievement through increasing vulnerability

16 of marginalized groups by failing to address the underlying root causes of vulnerability and poverty that

17 are related to political economy, power dynamics and vested interests more broadly, instead treating the

18 symptoms as the cause (Magnan et al., 2016; Ajibade and Egge, 2019; Schipper, 2020). For example,

19 evidence exists of flood defense measures through large scale infrastructure development leading to the

20 violent displacement of poor communities, forcibly resettling people in areas far from their employment

21 or pushing up land and housing costs without providing compensation (Fuso Nerini et al., 2018; Reckien

22 et al., 2018). Moreover, sectoral approaches to adaptation that fail to acknowledge the linkages between

23 SDGs can counter development efforts and generate further tradeoffs (Terry, 2009; Rasul and Sharma,

24 2016; von Stechow et al., 2016; Klinsky et al., 2017; Hallegatte et al., 2019).

25

26 The literature recommends a set of strategies for ensuring that adaptation actions are aligned with SDG

27 achievement and do not further perpetuate poverty and inequality. These include ensuring that

28 marginalized voices are central to adaptation decision-making, with participatory approaches that

29 empower and compensate affected communities (Moser and Ekstrom, 2011; Broto et al., 2015; Pelling

30 and Garschagen, 2019; Palermo and Hernandez, 2020). Gender mainstreaming and gender transformative

31 approaches within climate policies can also help ensure gender-sensitive design of adaptation projects,

32 with appropriate equity analyses of policy (Klinsky et al., 2017) decisions to identify the actual

33 implications of trade-offs for vulnerable groups (Beuchelt and Badstue, 2013; Alston, 2014; Bowen et al.,

34 2017; Fuso Nerini et al., 2018).

35

36 In addition, a substantial literature also argues for policy coherence measures that adopt whole-of-

37 government approaches and mainstream and nationalize SDG targets within national climate policies

38 (Nilsson et al., 2012; Le Blanc, 2015; Ari, 2017; Collste et al., 2017; Dzebo et al., 2017; Nilsson and

39 Weitz, 2019). Institutional coordination mechanisms that aim to break down silos between different

40 agencies and actors at the national level are suggested as beneficial for avoiding tradeoffs between

41 adaptation actions and SDGs (Mirzabaev et al., 2015; Howlett and Saguin, 2018; Scherer et al., 2018).

42 However, these need to be paired with an investigation of the deep-seated ideologies and vested interests

43 that are creating goal conflicts and negatively impacting marginalized groups to begin with (Purdon,

44 2014; Bocquillon, 2018). Ultimately, adaptation measures need to acknowledge and address the

45 underlying drivers that make certain groups particularly vulnerable, such as social disenfranchisement,

46 unequal power dynamics and historical legacies of colonialism and exploitation (Magnan et al., 2016;

47 Schipper, 2020)

48

49

50 Table Box 18.4.1: Examples of linkages between adaptation and the SDGs. For several key SDGs aligned with the

51 concept of CRD, the table below identifies evidence from the literature where adaptation policies and practices

52 contribute to achievement of the SDG as well as where they undermine achievement of the SDG.

    SDG             Evidence of adaptation contributing to       Evidence of adaptation undermining SDG

                                      SDG

    SDG 2: Zero Adaptation measures implemented by Some adaptation policies can increase land and food

    Hunger          smallholder farmers (e.g. adjustments in prices, negatively impacting smallholder farmers

                    farm operations timing, on-farm         (Fuso Nerini et al., 2018; Zavaleta et al., 2018;

                    diversification, soil-water management) Albizua et al., 2019)

    Do Not Cite, Quote or Distribute                 18-32                                       Total pages: 197
FINAL DRAFT                       Chapter 18                IPCC WGII Sixth Assessment Report

SDG 3: Good     exhibit higher levels of productivity and   Potential tradeoffs for food production through
Health and      technical efficiency in food production     adaptation actions within the water or energy sector, if
Wellbeing       (Bai et al., 2019; Sloat et al., 2020;      integrated approaches not taken (Howells et al., 2013;
                Khanal et al., 2021)                        FAO, 2014; Biswas and Tortajada, 2016)
SDG 6: Clean
Water and       Some climate smart agriculture              Societal measures beyond adaptation required to
Sanitation      measures (e.g. intercropping) can           address underlying causes of inequities that drive poor
                significantly increase yields and           health and well-being, including cuts in public
                contribute to zero hunger (Lipper et al.,   spending and neoliberalization and commodification
                2014; Arslan et al., 2015; Saj et al.,      of healthcare (Hall, 2020; Walsh and Dillard-Wright,
                2017)                                       2020)
                Increased resilience of societies and
                reduced vulnerability through               Potential tradeoffs for water security through
                investments in public health care and       adaptation actions within the food or energy sector, if
                access (Marmot, 2020; Mullins and           integrated approaches not taken (Howells et al., 2013;
                White, 2020)                                Rasul and Sharma, 2016; Mpandeli et al., 2018)

                Adaptation measures that leverage
                solidarity, equity and nature
                connectedness contribute to physical and
                psychological health and wellbeing
                (Gambrel and Cafaro, 2009; Capaldi et
                al., 2015; Soga and Gaston, 2016;
                Woiwode, 2020)
                Integrated water resources management
                as an adaptation strategy (Tan and Foo,
                2018; Sadoff et al., 2020)

                                                            Local, regional, or national "grabs" for water from
                                                            shared resources to with poorly defined property
                                                            rights (Olmstead, 2014)

SDG 11:         Vulnerability reducing adaptation           Need to ensure that adaptation measures understand
Sustainable     measures that aim to upgrade informal       how power dynamics and cultural norms shape urban
Cities and      settlements, create affordable housing      form and communities' vulnerability and adaptive
Communities     and protect populations living in disaster  capacity (Sanchez Rodriguez et al., 2018)
                prone areas (Major et al., 2018; Sanchez
                Rodriguez et al., 2018; Ajibade and         Risk of built infrastructure aiming to increase
                Egge, 2019)                                 resilience ignoring local population needs and
                                                            creating low-skilled jobs that concentrate land, capital
                                                            and resources in the hands of the elite (Ajibade and
                                                            Egge, 2019)

SDG 16:         Potential for adaptation projects to        Studies from Bangladesh, Cambodia and Nepal found
Peace, Justice  support livelihoods incomes and             that climate change adaptation-related policies and
and Strong      resource management, and thereby            projects were an underlying cause of natural resource-
Institutions    reduce tensions and the risk of conflicts   based conflicts, as well as land dispossession and
                (Matthew, 2014; Dresse et al., 2018;        exclusion, entrenchment of dependency relations, elite
                Barnett, 2019)                              capture, and inequity (Sovacool, 2018; Sultana et al.,
                                                            2019)

                                                            Adaptation projects can reinforce top-down
                                                            knowledge and decision-making processes,
                                                            asymmetric power relations and elite capture of
                                                            adaptation resources (Nightingale, 2017; Eriksen et
                                                            al., 2021b)

                                                                                  Need for conflict-sensitive adaptation approaches that

                                                                                  aim to `do no harm' (Babcicky, 2013; Ide, 2020)

1
2

Do Not Cite, Quote or Distribute  18-33                     Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 [END BOX 18.4 HERE]

 2

 3

 4 18.2.5.2 Mitigation

 5

 6 Mitigation entails greenhouse gas emissions reductions, avoidance, and removal and sequestration, as well as
 7 management of other climate forcing factors (WGIII AR6). There are numerous individual and system
 8 mitigation options throughout the economy and within human and natural systems (very high confidence)
 9 (Chapter 16; 18.5). Limiting global average warming has been found to reduce climate risks (IPCC, 2018a;
10 IPCC, 2019b), and limiting global average warming to any temperature level has also been found to be
11 associated with broad ranges of emissions pathways representing socioeconomic, technological, market,
12 physical uncertainties (very high confidence) (Rose and Scott, 2018; Rose and Scott, 2020). Pathways
13 consistent with limiting warming to 2癈 and below have been found to require significant deployment of
14 mitigation options spanning energy, land use, and societal transformation (WGIII AR6 Chapter 3 and
15 Chapter 4; 18.3). and substantial economic, energy, land use, policy, and societal transformation (WGIII
16 AR6 Chapter 3 and Chapter 4). Such emissions pathways would represent deviations from current trends that
17 raise issues about their feasibility and therefore plausibility (Rose and Scott, 2018; Rose and Scott, 2020).

18

19 The technical and economic challenge of limiting warming has been found to increase non-linearly with
20 greater ambition, fewer mitigation options, less than global cooperative policy designs, and delayed
21 mitigation action (WGIII AR6 Chapter 3; Table 18.2). Table 18.2 provides a high-level summary of pathway
22 characteristic ranges based on the WGIII AR6 assessment. Global pathways find large regional differences
23 in mitigation potential, as well as the degree of regional non-linearity with greater mitigation ambition.
24 These represent opportunities for mitigation, but how this effort and cost would be facilitated and distributed
25 respectively is a policy question.

26

27 Table 18.2 illustrates that greater climate ambition implies more aggressive emissions reductions in each
28 region, and earlier regional peaking of emissions (if they have not peaked to date). Near-term regional
29 emissions increases are possible, even for 1.5癈 compatible pathways, but significantly lower emissions than
30 today are shown in all regions by 2050. Increases in total regional energy consumption, as well as fossil
31 energy, are observed for many pathways, even in the most ambitious where energy consumption growth is
32 potentially slower compared to less ambitious pathways. By 2050, regional fossil energy declines, but is not
33 eliminated in any region. Regional growth in electricity use is substantial in all pathways, even the most
34 ambitious, with the growth continuing and accelerating with time and regional dependence on electricity
35 (share of total energy consumption) also growing significantly. The broad ranges are an indication of
36 uncertainty and risk for regional transitions, noting that full uncertainty is likely broader than what is
37 captured by emissions scenario databases (Rose and Scott, 2018; Rose and Scott, 2020). Among other things,
38 pathways commonly assume idealized climate policies with immediate implementation; and model
39 infeasibilities (i.e., models unable to solve) increase with climate ambition and pessimism about mitigation
40 technologies (e.g., Clarke et al., 2014; Bauer et al., 2018; Rogelj et al., 2018; Muratori et al., 2020),
41 highlighting the increasing challenge and potential for actual infeasibility with lower global warming targets.
42 Together, Table 18.2 provides insights into the increasingly demanding system and development transitions
43 associated with lower global warming levels, as well as some of the low-carbon transition uncertainties and
44 risks (see also Figure 18.5).

45

46 Past assessment has evaluated representative mitigation strategies in terms of economic, technological,
47 institutional, socio-cultural, environmental/ecological, and geophysical viability, as well as relationships to
48 sustainable development goals (de Coninck et al., 2018). The strategies assessment analysis has been
49 updated for AR6 (Cross-Chapter Box FEASIB). These assessments identify types of barriers that could
50 affect an option's feasibility. Among other things, this work finds that, other than public transport and non-
51 motorized transport, every other mitigation option evaluated had at least one feasibility dimension that
52 represented a barrier or obstacle. The barriers also imply that there are trade-offs in these feasibility
53 dimensions to consider. The assessment of mitigation option-sustainable development relationships identifies
54 related literature and derives aggregate characterizations. Concerns about the potential sustainable
55 development implications of some mitigation technologies may be motivation for precluding the use of some
56 mitigation options. For instance, the potential food security and environmental quality implications of
57 bioenergy have received significant attention in the literature (e.g., Smith et al., 2013). However,

Do Not Cite, Quote or Distribute  18-34       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 constraining or precluding the use of bioenergy without or with CCS could have significant implications for
 2 the cost of pursuing ambitious climate goals, and potentially the attainability of those goals (e.g., Clarke et
 3 al., 2014; Bauer et al., 2018; Rogelj et al., 2018; Muratori et al., 2020). Bioenergy is not unique in this
 4 regard. Social and sustainability concerns have also been raised about the large-scale deployment of many
 5 low-carbon technologies, e.g., REDD+, wind, solar, nuclear, fossil with CCS, and batteries. See WGIII
 6 Chapter 3 for examples of the potential implications of limiting or precluding different low-carbon
 7 technologies.

 8

 9 Overall, like with adaptation options, insights from this aggregate feasibility and sustainable development
10 mapping work are high-level and difficult to apply to a specific mitigation context. The feasibility, ranking,
11 and sustainable development implications of mitigation options, as well as the list of options themselves, for
12 a given location will vary from location-to-location, with different criteria and weighting of criteria that
13 reflect the relevant social priorities and differences in markets, technology options, and policies for
14 managing risks and trade-offs. Integrated evaluation of criteria and options is needed here as well, that
15 accounts for the relevant geographic context and interactions between options, systems, and implications.

16

17 Analyses of the potential implications of mitigation on sustainable development has various strands of
18 literature--studies exploring general greenhouse gas mitigation feedbacks to society, assessments of
19 mitigation implications on specific societal objectives other than climate, and literature evaluating mitigation
20 implications specifically for sustainable development objectives (WGIII AR6 Chapter 3, Chapter 4, Chapter
21 17). In general, mitigation alters development opportunities by constraining the emissions future society can
22 produce, which affects markets, resource allocation, economic structure, income distribution, consumers, and
23 the environment (besides climate) (very high confidence). Examples of general development feedbacks from
24 mitigation, include estimated price changes, macroeconomic costs, and low carbon energy and land system
25 transformations (e.g., WGIII AR6 Chapter 3 and Chapter 4) (Fisher et al., 2007; Clarke et al., 2014; Popp et
26 al., 2014; Rose et al., 2014; Weyant and Kriegler, 2014; Bauer et al., 2018; Rogelj et al., 2018). Examples of
27 mitigation implications for specific other variables of societal interest include evaluating potential effects on
28 air pollutant emissions, crop prices, water, and land use change (e.g., McCollum et al., 2018b; Roy et al.,
29 2018), while the literature evaluating mitigation implications specifically for sustainable development
30 objectives includes evaluations on energy access, food security, and income equality (e.g., Roy et al., 2018;
31 Arneth et al., 2019; Mbow et al., 2019). Proxy indicators are frequently used to represent whether there
32 might be implications for a sustainable development objective. For example, changes in energy prices are
33 used as a proxy for effects on energy security (e.g., Roy et al., 2018). This is common with aggregate
34 modelling studies, like those associated with global or regional emissions scenarios and energy systems.

35

36 Figure 18.5, derived from WGIII scenarios data, illustrates estimated relationships between mitigation and
37 various sustainable development proxy variables for different global regions. Figure 18.5 illustrates
38 synergies and trade-offs with mitigation, as well as regional heterogeneity, that can intensify with the level
39 of climate ambition--synergies in air pollutants, such as black carbon, NOx, and SO2; and trade-offs in
40 overall economic development, household consumption, food crop prices, and energy prices for electricity
41 and natural gas. For comparison, recent IPCC assessments also observed similar synergies and trade-offs but
42 did not directly make comparisons regarding overall development nor evaluate potential climates above 2癈
43 (Rogelj et al., 2018; Roy et al., 2018; Mbow et al., 2019). Regional non-linearity in the economic costs of
44 mitigation with greater climate ambition (i.e., costs rising at an increasing rate with lower warming goals)
45 can be significant within individual models (Rose and Scott, 2018; Rose and Scott, 2020). Figure 18.5 also
46 illustrates transition risks in the potential for significant synergistic and trade-off implications with, for
47 instance, potentially large regional commodity price implications and household consumption losses, as well
48 as more significant air pollution benefits. Note that the 1.5癈 results in Figure 18.5 (and Table 18.2) are
49 biased by model infeasibilities. Many models are unable to solve, especially with less optimistic
50 assumptions, resulting in small sample sizes and a different representation of models compared to the 2癈
51 and higher results.

52

53 Results like those in Figure 18.5 illustrate that mitigation-development trade-offs and balancing of societal
54 priorities are inevitable and need to be considered. For instance, Roy (2018) found that none of the 1.5癈 and
55 2癈 pathways assessed achieved all of the UN's Sustainable Develop Goals (SDGs). A newer literature is
56 developing evaluating the potential for managing SDG trade-offs. For instance, Roy et al. (2018) discuss the
57 potential for policies that address distributional implications, such as payments, food support, revenue

Do Not Cite, Quote or Distribute  18-35       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 recycling, as well as education, retraining, and technology outreach, subsidies, or prioritization. Recent
 2 studies have begun to estimate potential payments to offset trade-offs, such as related to food, water, and
 3 energy access (e.g., McCollum et al., 2018a). These analyses estimate investments to address specific trade-
 4 offs; however, with mitigation redirecting resources away from other productive activities, there is a need to
 5 also evaluate the aggregate economy-wide, distributional, and welfare effects, including the redistribution
 6 effects of managing sustainable development trade-offs.

 7

 8 There are a wide range of mitigation options and systems to consider, with assessment suggesting that a
 9 diverse portfolio is practical for pursing climate policy ambitions. However, local context will impact
10 mitigation choices, with unique sustainable development priorities, available mitigation options, sustainable
11 development synergies and trade-offs, and policy design and implementation possibilities.

12

13

Do Not Cite, Quote or Distribute  18-36       Total pages: 197
   FINAL DRAFT                                             Chapter 18              IPCC WGII Sixth Assessment Report

1 Table 18.2: Emissions pathway regional characteristics from WGIII scenarios database for pathways associated with different global warming levels (1.5癈, 2癈, 3癈, and 4癈).

2 Sample sizes: n = 13-15, 151-160, 66, and 34 emissions pathways for 1.5癈, 2癈, 3癈, and 4癈 global warming levels respectively. Sample size ranges for the same warming level

3 indicate that the sample size varies by variable due to differences in model reporting. Sample size varies by warming level due to model infeasibilities and differences in model

4 reporting.

                             Peak

   Variable                   global               Asia                Latin America             Middle East / Africa                  OECD              Reforming Economies
                             warming

                             to 2100

                             1.5癈                 2020                      2010 to 2030              2010 to 2030              2010 to 2020                  2015 to 2030

   Peak CO2 emissions        2癈             2015 to 2030                    2010 to 2035              2010 to 2030              2010 to 2020                  2015 to 2030

              year           3癈             2020 to 2080                    2010 to 2100              2030 to 2100              2010 to 2002                  2015 to 2100

                             4癈             2030 to 2100                    2010 to 2100              2070 to 2100              2010 to 2100                  2040 to 2100

                             Peak                  Asia                Latin America             Middle East / Africa                  OECD              Reforming Economies

   Variable                   global   2030                2050        2030                2050  2030                2050  2030                2050      2030                2050
                             warming
                              to 2100

                             1.5癈     -36 to 10%        -89 to -55%   -61 to 19%  -98 to 68%    -26 to 40%  -73 to -41% -56 to -24%    -96 to -78%      -42 to 14%  -95 to -48%

   Net CO2 emissions (%      2癈       -31 to 50%        -89 to -29%   -62 to 31%  -98 to -3%    -30 to 67%  -66 to 8%     -50 to -11%  -96 to -48%      -52 to 33% -105 to -27%

   from 2010)                3癈       10 to 50%         -5 to 69%     -58 to 16%  -132 to 50%   7 to 84%    37 to 158%    -44 to 2%    -69 to -12%      -18 to 34%  -35 to 41%

                             4癈       26 to 80%         18 to 205%    -49 to 26%  -41 to 36%    19 to 121%  78 to 225%    -30 to 8%         -55 to 5%   -13 to 36%  0 to 77%

                             1.5癈     9 to 57%          1 to 87%      18 to 68%   17 to 146%    31 to 57%   51 to 91%     -16 to 8%         -43 to 3%   -21 to 10%  -41 to 21%

   Energy consumption        2癈       17 to 91%         16 to 130%    3 to 72%    8 to 162%     18 to 82%   42 to 145%    -16 to 10%        -36 to 25%  -15 to 37%  -33 to 29%

   growth (% from 2010)      3癈       43 to 80%         70 to 129%    -9 to 74%   17 to 170%    21 to 82%   81 to 174%    -16 to 13%        -28 to 21%  -3 to 37%   -6 to 86%

                             4癈       47 to 109%        88 to 245%    20 to 65%   36 to 163%    47 to 95%   94 to 254%    -9 to 7%          -15 to 31%  -8 to 37%   -4 to 66%

                             1.5癈     -23 to 39%        -51 to 7%     -12 to 47%  -66 to 30%    -4 to 40%   -38 to -2%    -47 to -9%   -86 to -40%      -38 to 5%   -85 to -17%

   Fossil energy use growth  2癈       -33 to 66%        -73 to 18%    -20 to 65%  -78 to 63%    -6 to 71%   -78 to 61%    -47 to -8%   -78 to -28%      -51 to 31%  -84 to 18%

   (% from 2010              3癈       15 to 70%         29 to 103%    -20 to 65%  -10 to 124%   7 to 79%    31 to 158%    -37 to 3%         -61 to 3%   -24 to 32%  -26 to 43%

                             4癈       38 to 112%        39 to 264%    12 to 63%   24 to 176%    41 to 115% 103 to 301%    -26 to -5%        -45 to 10%  -14 to 29%  -5 to 66%

                             1.5癈     58 to 178% 141 to 463% 86 to 156% 275 to 430% 95 to 155% 296 to 791%                3 to 26%     32 to 103%       2 to 45%    45 to 173%

   Electricity consumption   2癈       41 to 232% 109 to 580% 11 to 156%           68 to 489%    27 to 172%  88 to 749%    -2 to 35%    16 to 143%       -8 to 112%  18 to 187%

   growth (% from 2010)      3癈       57 to 198% 126 to 472% 34 to 129% 140 to 364% 75 to 175% 260 to 600%                -3 to 39%    15 to 128%       3 to 112%   38 to 221%

                             4癈       107 to 243% 203 to 568% 49 to 127% 157 to 416% 87 to 200% 332 to 752%               10 to 33%         20 to 88%   36 to 83%   78 to 190%

     Electricity share of    1.5癈     -6 to 67%         12 to 166%    26 to 47%   61 to 181%    24 to 70%   100 to 258%   -2 to 21%    23 to 126%       -14 to 39%   9 to 145%
    energy consumption        2癈      -10 to 69%        2 to 156%     -13 to 79%  -1 to 161%    -9 to 72%   10 to 227%    -11 to 22%   11 to 121%       -18 to 57%  -11 to 143%
   growth (% from 2010)       3癈      -7 to 69%         5 to 134%     -9 to 79%   20 to 146%    -4 to 80%   42 to 149%    -12 to 33%                    -12 to 57%   6 to 100%
                              4癈      28 to 66%         40 to 120%    18 to 44%   46 to 95%     30 to 55%   87 to 142%     4 to 25%     7 to 87%        27 to 59%    43 to 98%
                                                                                                                                        13 to 69%

5

   Do Not Cite, Quote or Distribute                        18-37                                       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1

2 Figure 18.5: Implications of mitigation for different global mean temperature outcomes on various development and
3 sustainable development proxy variables. Example of 2050 global implications of mitigation for different global mean
4 temperature outcomes on various development and sustainable development proxy variables. Developed from the

Do Not Cite, Quote or Distribute  18-38       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 scenarios associated with (Bauer et al., 2018). Data sample sizes (not shown, but to be added) vary across temperature
 2 levels and variables due to model infeasibilities and model differences in reporting.

 3

 4

 5 18.2.5.3 Combining adaptation, mitigation, and sustainable development options

 6

 7 In practice, adaptation, mitigation, and sustainable development interventions are likely to be implemented
 8 in portfolio packages rather than as individual discrete options in isolation (high agreement, limited
 9 evidence). However, there is a dearth of literature estimating optimal portfolios of global adaptation and
10 mitigation strategies. This is not surprising given the geographic-specific nature of climate impacts and
11 adaptation and the information and computational complexity of representing that detail, as well as
12 mitigation options and interactions. There are, however, different literatures relevant to considering potential
13 combinations of adaptation, mitigation, and sustainable development.

14

15 At the most aggregate level, there is a long-standing literature exploring economically optimal global trade-
16 offs between climate risks and mitigation (e.g., Manne and Richels, 1992; Nordhaus, 2017; Rose, 2017), as
17 well as global stochastic analysis exploring global risk hedging for a small number of uncertainties (e.g.,
18 (Lemoine and Traeger, 2014). Recent work has found optimal global emissions and climate pathways to be
19 highly sensitive to uncertainties and plausible alternative assumptions, with uncertainties throughout the
20 causal chain from society to emissions to climate to climate damages shown to imply a wide range of
21 different possible economically optimal pathways (Rose, 2017). Among other things, this work identifies
22 assumptions consistent with limiting warming to different temperature levels. For example, the combination
23 of potential annual climate damages of 15% of global GDP at 4癈 of warming and a less sensitive climate
24 system were consistent with an economically efficient global pathway limiting warming to 2癈. In addition,
25 this work highlights the importance of characterizing and managing uncertainties. These types of global
26 aggregate analyses inform discussions regarding long-run global pathways and goals but are of limited value
27 to local near-term planning.

28

29 As discussed in Section 18.2.5.3.1, there are synergies and trade-offs mitigation, adaptation, and sustainable
30 development. For instance, the literature on the global cost-effectiveness of mitigation pathways provides
31 insights regarding aggregate synergies and trade-offs between mitigation and sustainable development (e.g.,
32 Figure 18.5). Furthermore, linkages between mitigation and adaptation options have been shown, such as
33 expected changes in energy demand due to climate change interacting with energy system development and
34 mitigation options, changes in future agricultural production practices to manage the risks of potential
35 changes in weather patterns affecting land based emissions and mitigation strategies, or mitigation strategies
36 placing additional demands on resources and markets which increases pressure on and costs for adaptation,
37 or ecosystem restoration that provides carbon sequestration and natural and managed ecosystem resiliency
38 benefits, but also could constrain mitigation and impact consumer welfare (WGIII AR6).

39

40 Non-linearities are an important consideration in evaluating risk management combinations. Non-linearities
41 have been estimated in global and regional mitigation costs and potential economic damages from climate
42 change (WGIII AR6 Chapter 3; (Clarke et al., 2014; Burke et al., 2015; Rose, 2017). Non-linear mitigation
43 costs mean increasingly higher costs for each additional incremental reduction in emissions (or incremental
44 reduction in global average temperature). Non-linear estimated economic climate damage means
45 increasingly higher damages for each additional incremental increase in climate change (e.g., global average
46 temperature) (very high confidence). Non-linearities are also suggested in estimated changes in key risks and
47 adaptation costs (Chapter 16, WGII sector and regional chapters). However, to date, they have not been as
48 explicitly characterized. These non-linearities imply non-linearities in climate risk management synergies
49 and trade-offs with sustainable development. Not only do trade-offs vary by climate level, as do synergies,
50 but they increase at an increasing rate and their relative importance can shift across climate levels (very high
51 confidence). Some of this is evident in results like those shown in Figure 18.5 for mitigation (keeping in
52 mind differences in sample sizes across temperature levels). Uncertainty about the degree of non-linearity in
53 mitigation, climate damages, key risks, and adaptation costs creates uncertainties in the strength of the trade-
54 offs and synergies, but also represents opportunities. For instance, additional mitigation options and more
55 economically efficient policy designs have been shown to reduce mitigation costs and the non-linearities in
56 mitigation costs (very high confidence) (WGIII AR6 Chapter 3). The same is true for adaptation options and
57 adaptation costs.

Do Not Cite, Quote or Distribute  18-39       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1

 2 Infeasibilities of mitigation and adaptation options (Section 18.4.2.2.1 and 18.4.2.2.2), as well as global
 3 pathways (WGIII AR6 Chapter 3), are also relevant to consideration of combinations of risk management
 4 options. Infeasibility of options implies higher costs and greater cost non-linearity due to fewer and/or more
 5 expensive options, while infeasibility of pathways bounds some of the uncertainty about the pathways
 6 relevant to decision-making and planning.

 7

 8 18.2.5.3.1 Trade-offs in adaptation, mitigation, and climate-resilient development
 9 Since AR5, a growing body of literature has emerged that frames adaptation processes as endogenous
10 socioeconomic dynamics, exogenous driving forces, and explicit decisions (Barnett et al., 2014; Maru et al.,
11 2014; Butler et al., 2016; Kingsborough et al., 2016; Werners et al., 2018). Central to this framing is a shift
12 away from viewing adaptation as discrete sets of options that are selected and implemented to manage risk,
13 to thinking about adaptation as a social process that evolves over time, includes multiple decision-points, and
14 requires dynamic adjustments in response to new information about climate risk, socioeconomic conditions,
15 and the value of potential adaptation responses (very high confidence) (Haasnoot et al., 2013; Wise et al.,
16 2016). This aligns adaptation with aspects of development thinking, including questions around the capacity
17 and agency of different actors to effect change, the governance of adaptation, and the contingent nature of
18 adaptation needs and effectiveness on the future evolution of society and climate change risk.

19

20 While ensuring development and adaptation produce synergies that allow for the achievement of sustainable
21 development is challenging, modelling exercises suggest that there are pathways where synergies among the
22 SDGs are realized (very high confidence) (Roy et al., 2018; Van Vuuren et al., 2019) (18.5), particularly if
23 longer time-horizons are used. These pathways require progress on multiple social, economic, technological,
24 institutional, and governance aspects of development including building human capacity, managing
25 consumption behavior, decarbonization of the global economy, improving food and water security,
26 modernizing cities and infrastructure, and innovations in science and technology (Van Vuuren et al., 2019)
27 (18.3). In addition, Olsson et al, (Olsson et al., 2014) and Roy et al. (2018) emphasize the importance of
28 integrating considerations for social justice and equity in the pursuit of sustainable development (Gupta and
29 Pouw, 2017).

30

31 The significant overlaps and linkages between development and adaptation practice and a lack of conceptual
32 clarity about adaptation pose a conundrum for scholars (e.g., Bassett and Fogelman, 2013; Webber, 2016),
33 who raise concerns that this potentially leads to trade-offs or mislabeling (Few et al., 2017). This framing of
34 adaptation and development can result in competition between attainment of sustainable development and
35 policies to reduce the impacts of climate change (Ribot, 2011). Such trade-offs are illustrated by (Moyer and
36 Bohl, 2019) who use a baseline development trajectory based on current trends to project progress on SDGs
37 by 2030. This work concluded that only marginal gains are likely to be achieved under that pathway over the
38 next decade (Barnes et al., 2019).

39

40 Emerging evidence also suggests that many adaptation-labelled strategies may exacerbate existing poverty
41 and vulnerability or introduce new inequalities, for example by affecting certain disadvantaged groups more
42 than others, even to the point of protecting the wealthy elite at the expense of the most vulnerable (Eriksen et
43 al., 2019). Pelling et al. (2016) find that adaptation has been conceived and implemented in such a manner
44 that most projects preserve rather than challenge the status quo. Specifically, the potential for knowledge and
45 the goals of adaptation to be contested by different actors and stakeholders and the need to sustain progress
46 over extended periods of time can constrain the ability to effectively implement actions that lead to
47 sustainable development outcomes that are protected from the impacts of climate change while also
48 delivering climate mitigation outcomes, that is, for climate resilient development (Bosomworth et al., 2017;
49 Bloemen et al., 2019). This creates the possibility for specific adaptation actions to result in outcomes that
50 undermine greenhouse gas mitigation and/or broader development goals (Fazey et al., 2016; Wise et al.,
51 2016; Magnan et al., 2020). For example, a study in Bangladesh revealed how local elites and donors used
52 adaptation projects as a lever to push vulnerable populations away from their agrarian livelihoods and into
53 uncertain urban wage labour (Paprocki, 2018). These types of outcomes are categorised as maladaptation,
54 interventions that increase rather than decrease vulnerability, and/or undermine or eradicate future
55 opportunities for adaptation and development (Barnett and O'Neill, 2010; Juhola et al., 2015; Magnan et al.,
56 2016; Antwi-Agyei et al., 2017; Schipper, 2020). This inadvertent impact on equity appears to
57 fundamentally contradict a benevolent understanding of transformative adaptation that also champions social

Do Not Cite, Quote or Distribute  18-40       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 justice (Patterson et al., 2018), thus posing long-term maladaptation in opposition to transformative
 2 adaptation (Magnan et al., 2020).

 3

 4 Similarly, mitigation efforts, while reducing emissions, can also increase climate impacts vulnerability and
 5 undermine adaptation efforts. The same can be said for some poverty alleviation and sustainable
 6 development efforts that increase vulnerability for specific segments of the population. For example, in
 7 Central America, an evaluation of twelve rural renewable energy projects (either for CDM, early warning
 8 systems or rural electrification goals) found that some mitigation and poverty alleviation projects increased
 9 vulnerability to families--by excluding them, not adhering to local safety and quality codes and standards, or
10 significantly altering community power dynamics and contributing to conflict (Ley, 2017; Ley et al., 2020).

11

12 Synergies between adaptation, mitigation and sustainable development might be promoted by prioritizing
13 those CRD strategies most likely to generate synergies (very high confidence) (Roy et al., 2018; Karlsson et
14 al., 2020). This could include focusing on poverty alleviation that improves adaptive capacity (e.g., Kaya and
15 Chinsamy, 2016; Kuper et al., 2017; Ley, 2017; S醤chez and Izzo, 2017; Staczuk-Galwiaczek et al., 2018;
16 Ley et al., 2020); renewable energy systems that improve water management and preservation of river
17 ecological integrity (e.g., Berga, 2016; Rasul and Sharma, 2016); or internalizing positive externalities, such
18 as subsidies for mitigation options thought to also improve water use efficiency (e.g., Roy et al., 2018).
19 Similarly, trade-offs might be managed by prioritizing strategies such as disqualifying mitigation options
20 thought to have negative social implications (Section 18.2.5.3.1), internalizing externalities, such as placing
21 a fee or constraint on a negative externality or related activity (e.g., WGIII AR6 Chapter 13) (Bistline and
22 Rose, 2018), or using complementary policies, such as transfer payments to offset negative mitigation,
23 adaptation, or sustainable development strategy implications (very high confidence) (e.g., McCollum et al.,
24 2018b). Roy et al. (2018) discusses the latter, noting, for instance, the possibility of complementary
25 sustainable development payments to avoid global energy access, food security, and clean water trade-offs.

26

27 SR1.5 and AR6 assessments of system transitions also find opportunities for synergies and managing trade-
28 offs (18.3; Cross-Chapter Box FEASIB). Within each system, mitigation and adaptation options are assessed
29 for their specific benefits and the impacts they can have on one another, as well as with sustainable
30 development. For example, within energy system transitions, the three adaptation options (power
31 infrastructure resilience, reliability of power systems, efficient water use management) have strong synergies
32 with mitigation. While not all mitigation options have strong synergies, the trade-offs can be managed when
33 adaptation and sustainable development goals are also considered. Under land and other ecosystems system
34 transitions, the main trade-off is the competition for land-use between potential alternative uses, e.g.,
35 sustainable agriculture, afforestation/reforestation, purpose-grown biomass for energy. On the other hand,
36 assessment of urban and infrastructure system transitions finds mainly synergies between mitigation and
37 adaptation options with trade-offs that are considered manageable, and there is growing evidence of rural
38 landscape infrastructure benefits to adaptation.

39

40 Overall, this literature is relatively new and still developing. It highlights the importance of sets of societal
41 priorities and policy design. However, it is not well developed in terms of joint optimization of multiple
42 priorities, evaluating alternative mechanisms and shifts in trade-offs, and evaluating redistribution
43 implications with transfers.

44

45 18.2.5.3.2 Risk management combinations with lower to higher climate change
46 The different strands of literature discussed above can be integrated to help inform thinking about
47 combinations of approaches to risk management. Globally, low climate change projections, versus higher
48 climate change projections, imply greater mitigation, lower climate risks, and less adaptation. This implies
49 greater mitigation trade-offs in terms of overall economic development, food crop prices, energy prices, and
50 overall household consumption, but lower climate risk, with sustainable development synergies like human
51 health and lower adaptation trade-offs, and an uneven distribution of effects (very high confidence) (Roy et
52 al., 2018).

53

54 Sustainable development considerations could be used to prioritize mitigation options, but as noted earlier
55 there are trade-offs, with a potentially significant impact on the economic cost of mitigation, as well as a
56 potential trade-off in terms of the climate outcomes that are still viable (WGIII AR6 Chapter 3). For
57 instance, all of the 1.5癈 scenarios used in IPCC (2018a) deploy carbon dioxide removal technologies

Do Not Cite, Quote or Distribute  18-41       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 (Rogelj et al., 2018). Without these technologies, most models cannot generate pathways that limit warming
 2 to 1.5癈, and those that do adopt strong assumptions about global policy development and socioeconomic
 3 changes. Sustainable development might also affect the design of policies by prioritizing specific sustainable
 4 development objectives. However, there are trade-offs here as well, with costs and the distribution of costs
 5 varying with alternative policy designs. For instance, prioritizing air quality has climate co-benefits but does
 6 not ensure the lowest cost climate strategy (Arneth et al., 2009; Kandlikar et al., 2009). Similarly,
 7 prioritizing land protection has a variety of co-benefits but could increase food prices significantly, as well as
 8 the overall cost of climate mitigation (IPCC, 2019b). In this context with lower climate risk and adaptation
 9 levels and larger mitigation effort, managing mitigation trade-offs could be a sustainable development
10 priority. Furthermore, sustainable development could also be tailored to facilitate adaptation as well as
11 manage mitigation costs.

12

13 Globally, high climate change projections imply lower mitigation effort, higher climate risks, and greater
14 adaptation. This implies lower mitigation trade-offs, but greater climate risk with greater demand of
15 adaptation and potential for trade-offs in terms of competing sustainable development priorities. Sustainable
16 development considerations could affect adaptation options. For instance, constraining options such as
17 relocation or facilitating adaptation capacity and community resilience. Sustainable development might also
18 be tailored to affect the climate outcome by shaping the development of emissions. In this context with
19 greater climate risk and adaptation levels and less mitigation effort, facilitating adaptation and managing
20 adaptation costs and trade-offs could be a sustainable development priority.

21

22 Locally, there are many qualitative similarities to the global perspective in thinking about risk management
23 combinations across lower versus higher climates. However, there is one very important difference. Local
24 decision makers are confronted with uncertainty about what others will do beyond their local jurisdiction.
25 With future climate a function of the sum of global decisions, sustainable development planning needs to
26 consider the possibility of more and less emissions reduction action globally and the potential associated
27 climates. This implies the need for sustainable development to manage for the possibility of higher climates
28 by further facilitating adaptation and managing adaptation trade-offs. Prioritizing sustainable development
29 locally is also supported by the insight that the impacts on poverty depend at least as much or more on
30 development than on the level of climate change (very high confidence) (Wiebe et al., 2015; Hallegatte and
31 Rozenberg, 2017).

32

33 There is nothing in the current literature to suggest that CRD is necessarily associated with a specific climate
34 outcome, like limiting global average warming 1.5癈 or 2癈, or a specific pathway. Instead, there are many
35 possible pathways for climate-resilient development (medium agreement, limited evidence) (e.g., David
36 T郻ara et al., 2018; O'Brien, 2018). The current literature suggests that different mixes of adaptation and
37 mitigation strategies, and sustainable development and trade-off management priorities, measures, and
38 reallocations (Section 18.5.3.1), will be appropriate for different expected climates and locations (18.1.2);
39 while trade-offs between climates will be dictated by relative non-linearities, feasibilities, shifts in priorities,
40 and trade-off and reallocation options across future climates.

41

42 Finally, it is important to note that there is currently limited information available regarding the following:
43 (1) local implications of 1.5癈 versus warmer futures with respect to avoided impacts and sustainable
44 development implications and interactions and applying global conclusions to local, national, and regional
45 settings can be misleading, (2) local context-specific synergies and trade-offs with respect to adaptation,
46 mitigation, and sustainable development for 1.5癈 futures, and (3) standard indicators for monitoring factors
47 related to CRD (Roy et al., 2018).

48

49

50 18.3 Transitions to Climate Resilient Development

51

52 A key finding emerging from the IPCC SR1.5 is the critical role that system transitions play in enabling
53 mitigation pathways consistent with a 1.5癈 or less world (IPCC, 2018a; IPCC, 2019b).Such transitions are
54 similarly critical for the broader pursuit of climate-resilient development, and the various AR6 special
55 reports as well as subsequent literature provide new evidence of why such transitions are needed for CRD, as
56 well as both the opportunities for accelerating system transitions and their limitations for delivering on the
57 goals of CRD.

Do Not Cite, Quote or Distribute  18-42       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1

 2 18.3.1 System Transitions as a Foundation for Climate Resilient Development

 3

 4 In the AR6, system transitions are defined as "the process of changing (the system in focus) from one state or
 5 condition to another in a given period of time" (IPCC, 2018a; IPCC, 2019b). In the climate change solution
 6 space, system transitions represent an important mechanism for linking and enabling mitigation, adaptation,
 7 and sustainable development options and actions (very high confidence). SR1.5C identified the need for
 8 rapid and far-reaching transitions in four systems � energy, land and terrestrial ecosystems, urban and
 9 infrastructure, and industrial systems (IPCC, 2018b; IPCC, 2018a) (1.5.1 and 18.1). The SRCCL expanded
10 on this with a focus on terrestrial systems, while SROCC added additional evidence from ocean and
11 cryosphere systems. This section assesses the four system transitions discussed in the SR1.5C assessment in
12 the context of CRD, while also extending the assessment to consider societal transitions as a cross-cutting,
13 fifth transition important for climate-resilient development. Literature to support this assessment is also
14 drawn from AR6 regional and sectoral chapters, which is synthesized later in this chapter (18.5).

15

16 As discussed in Box 18.3 (H鰈scher et al., 2018), system transitions are linked to system transformation,
17 which is defined as "a change in the fundamental attributes of a system including altered goals or values"
18 (Figure 18.1) (IPCC, 2018a). In a systems context, transitions focus on `complex adaptive systems; social,
19 institutional and technological change in societal sub-systems', while transformations are "large scale
20 societal change processes ... involving social-ecological interactions" (IPCC, 2018a) (Box 18.1). Although
21 system transitions are often identified in the literature as being necessary processes for large-scale
22 transformations (Roggema et al., 2012; H鰈scher et al., 2018), thereby making them a core enabler of CRD.
23 Yet, they are not necessarily transformative in themselves.

24

25 18.3.1.1 Energy Systems
26 Recent observed changes in global energy systems include continued growth in energy demand, led by
27 increased demand for electricity by industry and buildings (very high confidence) (AR6 WGIII Chapter 2).
28 Growth in energy demand has also been driven by increased demand for industrial products, materials,
29 building energy services, floor space, and all modes of transportation. This growth in demand, however, has
30 been moderated by improvements in energy efficiency in industry, buildings, and transportation sectors (very
31 high confidence) (AR6 WGIII Chapter 2). There is also a trend of moving away from coal towards cleaner
32 fuels, due to lower natural gas prices and lower cost renewable technologies, and structural changes away
33 from more energy-intensive industry.

34

35 Features of sustainable development such as enhanced energy access, energy security, reductions in air
36 pollution, and economic growth continue to be the dominant influence on the evolution of energy systems
37 and decision-making regarding energy investments and portfolios (very high confidence) (WGIII AR6
38 Chapter 6). To date, climate policy has been comparatively less influential in driving energy transitions
39 globally. Yet, there are examples at the local, regional, and national level of policy incentivizing rapid
40 changes in energy systems (very high confidence) (WGIII AR6 Chapter 6). Many sustainable development
41 priorities have co-benefits in terms of climate mitigation, such as air pollution and conservation policies
42 reducing short-lived climate forcers and sequestering carbon respectively, as well adaptation benefits, such
43 as improved energy access and environmental quality enhancing adaptive capacity (very high confidence)
44 (WGIII AR6 Chapter 6) (de Coninck et al., 2018). Alternatively, sustainable development projects can have
45 negative climate implications with, for instance, hydroelectric projects shut down by droughts or floods
46 resulting in greater use of bunker and fuel oil, as well as natural gas.

47

48 In addition to sustainable development priorities driving change in energy systems, observed energy system
49 trends have implications for sustainable development (e.g., IEA et al., 2019). Observed changes in energy
50 system size, rate of growth, composition and operations impact energy access, equity, environmental quality
51 and wellbeing, with both synergies and trade-offs, including recent improvements in global access to
52 affordable, reliable, and modern energy services. For instance, in some countries, such as the United States,
53 there has been a significant shift away from coal as a fuel source for electricity generation in favor of natural
54 gas. More recently, however, renewables have emerged as the dominant form of new electricity generation
55 (Gielen et al., 2019). Similarly, for energy access in developing countries, renewable energy or hybrid
56 distributed generation systems are increasingly being prioritized due to challenges associated with access,

Do Not Cite, Quote or Distribute  18-43       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 costs and environmental impacts from traditional fossil fuel-based energy technologies (Mulugetta et al.,
 2 2019).

 3

 4 Energy systems have been a historical driver of climate change, but are also adversely affected by climate
 5 change impacts, including short-term shocks and stressors from extreme weather as well as long-term shifts
 6 in climatic conditions (very high confidence). The potential for such factors is often incorporated into local
 7 system designs, operations, and response strategies. There have been changes in observed weather and
 8 extreme event hazards for the energy system, but to date many are not attributable solely to anthropogenic
 9 climate change (USGCRP, 2017; IPCC, 2021a). Nevertheless, with observed extremes shifting outside of
10 what has been observed historically, existing design criteria and operations may not be optimal for future
11 climate conditions and contingencies (Chapter 16; sectoral and regional chapters). Overall, there is limited
12 historical evidence on the efficacy of adaptation responses in reducing vulnerability of energy systems (high
13 agreement, limited evidence). However, sustainable development trends, such as improving incomes,
14 reducing poverty, and improving health and education have reduced vulnerability (Chapter 16), and
15 improvements in system resiliency to extreme weather events and more efficient water management have
16 occurred that have synergies with adaptation and sustainable development in general.

17

18 Available literature indicates that greenhouse gas emissions reductions have been achieved in response to
19 climate actions including financial incentives to promote renewable energy, carbon taxes and emissions
20 trading, removal of fossil fuel subsidies, and promotion of energy efficiency standards (very high
21 confidence) (WGIII AR6 Chapter 6). Such policies tend to lead to a lower carbon intensity of GDP, due to
22 structural changes in the use of energy and the adoption of new energy technologies. However, other drivers
23 of change are also present and thus ongoing energy transitions and their future evolution are a response to
24 both climatic and non-climatic considerations, with broader sustainable development priorities being a
25 significant driver of change (see WGIII AR6 Chapter 6).

26

27 18.3.1.2 Urban and infrastructure systems

28

29 Urban areas their associated infrastructure are critical targets for CRD processes. This is a function of urban
30 areas being the dominant settlement pattern with over 55% of the global population living in cities (World
31 Bank, 2021). As a consequence, urban areas are also the focal point for energy use, land use change, and
32 consumption of natural resources, thereby making them responsible for an estimated 70% of global CO2
33 emissions (Johansson et al., 2012; Ribeiro et al., 2019). The trend toward increasing urbanization is
34 anticipated to create both challenges and opportunities for sustainable development, as well as climate action
35 (G黱eralp et al., 2017; Li et al., 2019a).

36

37 The built environment is increasingly exposed to climate stresses and more frequent co-occurrences of
38 climate shocks than in the past. This has the potential to increase rates of building and infrastructure
39 degradation, increase damage from extreme weather events. The existing adaptation gaps and everyday risks
40 within many cities, particularly those of the global South, combined with escalating risk from climate
41 change, makes rapid progress in enhancing urban resilience a high priority for CRD (Pelling et al., 2018;
42 Davidson et al., 2019; Lenzholzer et al., 2020). Strategic investments in disaster risk reduction, including
43 climate-resilient green infrastructure, updated building codes, and land use planning can provide significant
44 long-term cost savings and social benefits. Moreover, evaluating the relative merits of "fail safe" versus
45 "safe to fail" approaches to infrastructure planning can help to identify more design principles that are more
46 robust to the uncertainties of climate change and urbanization (Kim et al., 2017a; Kim et al., 2019).

47

48 Much of the literature on urban resilience and sustainability focuses on addressing discrete challenges for
49 urban infrastructure sub-systems. Climate change has the potential to enhance stress on lifeline infrastructure
50 services such as the provision of electricity, water and wastewater, communications, and transportation �
51 sub-systems which often underdeveloped in many regions of the world (Arku and Marais, 2021; Sitas et al.,
52 2021). For example, a warming and more variable climate can increase stress on electricity grids by reducing
53 transmission efficiency, increasing cooling demand requirements, and by increasing exposure to climate
54 shocks such as heat waves, floods, and storms (Bartos and Chester, 2015; Auffhammer et al., 2017; Perera et
55 al., 2020). Accordingly a significant focus on the energy transition is on achieving the dual goals of reducing
56 the carbon footprint of energy while also increasing resilience of energy supply to current and future threats.

Do Not Cite, Quote or Distribute  18-44       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 For example, renewable energy generation and storage technologies that modular and distributed and
 2 provide enhanced resilience to shocks and stresses from climate change (Venema and Temmer, 2017a).

 3

 4 Similarly, building and maintaining urban water systems that are resilient to climate shocks requires
 5 significant changes in water demand, infrastructure, and management. Enhancing redundancy in water
 6 supply and the flexibility to shift between surface and groundwater options aids adaptation. Decentralized
 7 water supply and sanitation options are now feasible and can provide greater resilience than most centralized
 8 systems (Parry, 2017), provided they have adequate supply (Leigh and Lee, 2019; Rabaey et al., 2020).
 9 Water conservation and green infrastructure options for stormwater management are proven approaches for
10 reducing climate risks (Venema and Temmer, 2017b), with adaptation and mitigation co-benefits. Water
11 demand management and rainwater harvesting contribute to climate change mitigation and increase adaptive
12 capacity by increasing resilience to climate change impacts such as drought and flooding (Paton et al., 2014;
13 Berry et al., 2015). In addition, they can contribute to restoring urban ecosystems that offer multiple
14 ecosystem services to citizens (Berry et al., 2015) (see WGIII AR6 Chapter 8). The context-appropriate
15 development of green spaces, protecting ecosystem services and developing nature-based solutions, can
16 increase the set of available urban adaptation options (IPCC, 2018b), while creating opportunities for more
17 complex and dynamic approaches to urban water management (Franco-Torres et al., 2020). For example, the
18 Netherlands' `Room for the River' policy focuses on ont only achieving higher flood resilience, but also
19 improving the quality of riverine areas for human and ecological wellbeing (Busscher et al., 2019).

20

21 An overarching focus of urban sustainability is the reversal of long-standing trends of ecosystem
22 fragmentation and degradation that have resulted in growing separation between human and natural systems
23 within urban environments (IPBES, 2019) (see WGIII AR6 Chapter 8). Urban ecosystems and the
24 integration of nature-based solutions and green infrastructure into urban areas can yield benefits that
25 facilitate achievement of the SDGs. There has been growing recognition of urban ecosystems as social,
26 cultural, and economic assets that can support economic development while also enhancing resilience to
27 extreme weather events and improving air and water quality (Shaneyfelt et al., 2017; Matos et al., 2019).
28 Investing in urban ecosystems and green infrastructure can provide lower-cost solutions to multiple urban
29 development challenges when compared to traditional infrastructure systems (Terton, 2017). Relatedly,
30 agriculture, while largely a rural system, is increasingly expanding within urban areas. Urban agriculture
31 enables citizens to fulfil some of their food needs, improving urban resilience to food shortages, enhancing
32 biodiversity, and increasing coping capacity during disasters (Demuzere et al., 2014; Clucas et al., 2018) (see
33 WGIII AR6 Chapter 8). Strengthening urban agroecosystems therefore increases resilience to supply shocks
34 from climate change impacts and can contribute to community cohesion (Temmer, 2017a).

35

36 Overall, the discourse in the literature regarding the future of cities emphasizes the importance of viewing
37 cities as more than just their physical infrastructure that can be made more resilient through engineering
38 solutions (Davidson et al., 2019). Rather, urban areas are increasingly conceptualized as complex
39 socioecological or sociotechnical systems (very high confidence) (Patorniti et al., 2017; Patorniti et al., 2018;
40 Visvizi et al., 2018; Savaget et al., 2019). Such frameworks integrate physical, cyber, social, and ecological
41 elements of cities in pursuit of resilience and sustainability transitions, and they recognize the role of
42 governance and engagement processes as being central to system change (Temmer, 2017b). Nevertheless,
43 some authors have cautioned that urban transitions will be associated with synergies as well as trade-offs
44 with respect to sustainable development (very high confidence) (Maes et al., 2019; Sharifi, 2020).

45

46

47 [START BOX 18.5 HERE]

48

49 Box 18.5: The Implications of the Belt and Road Initiative (BRI) for Climate Resilient Development

50

51 In 2013, Chinese President Xi Jinping announced plans for a grand transcontinental infrastructure initiative.
52 China would work with partner countries under two programs termed the Silk Road Economic Belt and the
53 21st Century Maritime Silk Road. Together, these have come to be known as the Belt and Road Initiative
54 (BRI). Set to encompass 4.4 billion people and a cumulative GDP of around $21 trillion, the BRI has been
55 implemented in over 120 countries with wide infrastructure funding gaps, as exemplified by the China-
56 Myanmar Gas Pipeline, Gwadar Port in Pakistan, Trans-Mongolian Railway, China Belarus Industrial Park,
57 and urban rehabilitation in Ethiopia. Its stated objectives even extend beyond infrastructure connectivity to

Do Not Cite, Quote or Distribute  18-45       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 include trade promotion, financial integration, policy coordination and cultural dialogue. Having been written
 2 into the Communist Party's constitution in 2017, the BRI will be China's flagship international development
 3 strategy for years to come.

 4

 5 The 126 countries participating in the BRI account for 23% of global GDP, but also 28% of global carbon
 6 emissions (PBCSF, 2019). By 2050, even based on an optimistic scenario, the total carbon emission by these
 7 countries will be 17% higher than what would be allowed under a 2癈 carbon budget (Duan et al., 2018). The
 8 BRI covers regions with high reserve of carbon-based fuels and could have significant impact on global energy
 9 consumption and carbon emission patterns. For example, according to the EIA statistics, the proven reserves
10 of oil, natural gas, and coal in nations under the BRI make up 58.8%, 79.9%, and 54.0% of the world's total
11 (China Meteorological Administration, 2019).

12

13 Meanwhile, countries along the BRI are highly vulnerable to the impact of climate change, spanning highly
14 diverse climate zones with fragile ecological conditions. Currently, many of the regions have a low level of
15 infrastructure development and high population densities (The People's Republic of China, 2017). Changes in
16 temperature, precipitation, vegetation and hydrological conditions could in turn pose threats to the
17 development and operation of infrastructure projects in these regions. Given the scope and scale of the BRI, a
18 key question is whether it will incentivize continued exploitation of available fossil fuel resources or provide
19 the innovation and economic development needed to transition participating nations to more resilient and less
20 carbon-intensive economies.

21

22 BRI and its commitment to climate resilient development (CRD)

23

24 Recognizing these feedbacks between the BRI and climate change, the Chinese government, included climate
25 change in developing the key guiding documents on BRI development in 2015. These include "taking into
26 consideration the impact of climate change, strengthening exchange and cooperation with countries along the
27 Belt and Road, leveraging the support and guarantee function of Chinese meteorological departments in
28 promoting the BRI" (NDRC, 2015). The second BRI Forum held in 2019 reiterated the importance of green
29 development "as the foundation of the BRI" and promoted green infrastructure development and green
30 investment, in addition to plans for increasing capacity in response to climate change, promoting low-carbon
31 infrastructure, energy source, climate-related disaster alarm system, climate finance integration, as well as
32 low-carbon technology development.

33

34 The Chinese Meteorological Administration, the governmental agency responsible for climate change related
35 issues, responded to BRI official guidelines by establishing BRI integrated meteorological service system and
36 proposed meteorological development plan 2017-2025 (China Meteorological Administration, 2019), which
37 includes policy coordination on climate change, promoting intergovernmental cooperation, completing BRI
38 disaster prevention and relief mechanisms, strengthening climate change support capacity, enhancing
39 prediction and evaluation capacity related with climate change (China Meteorological Administration, 2019).
40 China has established South-South cooperation in support of other countries to mitigate climate change. Efforts
41 have been made to promote joint research with countries along the BRI on regional climate change, climate
42 change prediction, and develop products in response to climate conditions in different regions.

43

44 The China Clean Development Mechanism Fund (CCDMF) is a national climate fund that supports low carbon
45 growth and climate resilience in China (UNFCCC, 2017). More than USD 81 million in grants committed to
46 support over 200 projects. A combination of funding enterprises, mobilizing market capital and achieving
47 verified emission reduction effects contributes to a direct reduction of over seven million tons of CO2
48 equivalent. Government representatives from Brazil, Vietnam, and Cambodia have already visited CCDMF to
49 learn more about this type of climate financing.

50

51 Trade-offs between BRI and CRD
52 Despite the implementation of such financing mechanisms for low-carbon development, their net effect is not
53 necessarily sufficient to offset the carbon footprint generated by overseas fossil fuel projects funded or
54 financed by China. As such, BRI stakeholders must navigate a number of trade-offs among different objectives
55 of the initiative.

56

Do Not Cite, Quote or Distribute  18-46       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 For the Chinese government and state-owned enterprises, an immediate trade-off is that between the short-
 2 term profits gained through carbon-intensive infrastructure investments overseas and long-term sustainable
 3 development with the introduction of low-carbon technology in infrastructure development. On one hand, the
 4 energy solutions that China proposes tend to involve carbon-intensive infrastructures such as coal factories,
 5 which increases carbon emissions of these countries. But at the same time, China also provides climate finance
 6 for these countries in support of renewable energy projects such as hydropower projects and solar panel
 7 production facilities.

 8

 9 For the governments and people hosting BRI projects, the tradeoff is between short-term economic prosperity
10 and long-term sustainable development. Infrastructure development driven by carbon-intensive technologies
11 are cheaper and more consistent for developing countries (for example, electricity generated through coal-
12 based power plants is more consistent than that generated through hydropower stations), which is conducive
13 to more rapid industrialization of these countries, generating immediate urbanization and economic prosperity.
14 Yet the industrialization process would exacerbate carbon emission and accelerate the climate change process,
15 with long-term impact on food security, livelihood, migration, water demand, disease control, posing potential
16 hazards to sustainable development in these regions.

17

18 Winners and losers in incorporating CRD into BRI development

19

20 An emphasis on CRD within the BRI could create a number of opportunities for sustainable development.
21 For example, adherence to CRD principles of low-carbon development would incentive growth of renewable
22 energy, clean technologies, thereby growing the global market for such goods and services. This could have
23 significant benefits for developing nations of the BRI in terms of enabling sustainability transitions that
24 might otherwise not be feasible. However, a CRD orientation of the BRI would also have consequences for
25 fossil fuel and carbon-intensive industries. This could affect both private and state-owned enterprises in BRI
26 nations resulting in stranded assets, loss of some forms of employment.

27

28 [END BOX 18.5 HERE]

29

30

31 18.3.1.3 Land, Oceans, and Ecosystems

32

33 Land, oceans, and terrestrial ecosystems are in transition globally, with anthropogenic factors including
34 climate change being a major driving force (very high confidence) (IPBES, 2019) (Box 6). Seventy-five per
35 cent of the land surface has been significantly altered, 66 percent of the ocean area is experiencing increasing
36 cumulative impacts, and over 85 percent of wetland areas have been lost (IPBES, 2019). Since 1970, only
37 four out of eighteen recognized ecosystem services assessed have improved in their functioning: agricultural
38 production, fish harvest, bioenergy production and material harvests. The other 14 ecosystem services have
39 declined (IPBES, 2019), raising concerns about the capacity of ecosystems and their services to support
40 sustainable and climate-resilient development.

41

42 Given the pressures on land, oceans, and ecosystems, enhancing resilience to climate change and other
43 pressures of human development is a core priority of transition in these systems. Yet, there are a few
44 recorded initiatives that provide evidence of successful improvement in ecosystem resilience (high
45 agreement, limited evidence). Similarly, although there is significant evidence that a broad range of
46 adaptation initiatives have been pursued across global regions and sectors, including a rapid expansion of
47 nature- or ecosystem-based solutions (Mainali et al., 2020), there is limited evidence of how these planned
48 climate adaptation efforts have contributed to enhanced ecosystem resilience. Additional research is
49 necessary to evaluate these efforts in terms of their performance and also to identify mechanisms for scaling
50 them up in different contexts. As an example, Paik (Paik et al., 2020) record the increased diffusion of salt
51 tolerant rice varieties in the Mekong River Delta, which is at risk of sea-level rise and an associated saline
52 intrusion. This is a low-cost adaption to saline ingress, that increases food productivity and reduces the risk
53 of outmigration for this vulnerable agricultural region.

54

55 Evidence of the interactions between ecosystems and resilience come from a range of sources including both
56 regional and sectoral examples (Box 18.2; Tables 18.7�18.8. For example, regional examples suggest that
57 the use of land to produce biofuels could increase the resilience of production systems and address

Do Not Cite, Quote or Distribute  18-47       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 mitigation needs (Box 2.2). Nevertheless, the potential of BECCS to induce maladaptation needs deeper
 2 analysis (Hoegh-Guldberg et al., 2019). Climate Smart Forestry (CSF) in Europe provides an example of the
 3 use of sustainable forest management to unlock the EU's forest sector potential (Nabuurs et al., 2017). This
 4 is in response to diverse climate impacts ranging from pressure on spruce stocks in Norway and the Baltics,
 5 on regional biodiversity in the Mediterranean region, and the opportunity to use afforestation and
 6 reforestation to store carbon in forests (Nabuurs et al., 2019). CSF considers the full value chain from forest
 7 to wood products and energy and uses a wide range of measures to provide positive incentives to firmly
 8 integrate climate objectives into the forestry sector. CSF has three main objectives; (i) reducing and/or
 9 removing greenhouse gas emissions; (ii) adapting and building forest resilience to climate change; and (iii)
10 sustainably increasing forest productivity and incomes (Verkerk et al., 2020).

11

12 Other solutions focus on specific subsectors. Mutually supportive climate and land policies have the
13 potential to save resources, amplify social resilience, support ecological restoration, and foster engagement
14 and collaboration between multiple stakeholders. (IPCC, 2019f, C.1). Land-based solutions can combat
15 desertification in specific contexts: water harvesting and micro-irrigation, restoring degraded lands using
16 drought-resilient ecologically appropriate plants, agroforestry, and other agroecological and ecosystem-based
17 adaptation practices (IPCC, 2019f, B.4.1). Reducing dust and sand storms and sand dune movement can
18 lessen the negative effects of wind erosion and improve air quality and health. Depending on water
19 availability and soil conditions, afforestation, tree planting and ecosystem restoration programs, using native
20 and other climate resilient tree species with low water needs, can reduce sand storms, avert wind erosion,
21 and contribute to carbon sinks, while improving micro-climates, soil nutrients and water retention (IPCC,
22 2019f, B.4.2).

23

24 Coastal blue carbon ecosystems, such as mangroves, salt marshes and seagrasses, can help reduce the risks
25 and impacts of climate change, with multiple co-benefits. Over 150 countries contain at least one of these
26 coastal blue carbon ecosystems and over 70 contain all three. Successful implementation of measures of
27 carbon storage in coastal ecosystems could assist several countries in achieving a balance between emissions
28 and removal of greenhouse gases. Carbon storage in marine habitats can be up to 1,000 tC ha�1, higher than
29 most terrestrial ecosystems. Conservation of these habitats would also sustain a wide range of ecosystem
30 services, assist with climate adaptation by improving critical habitats for biodiversity, enhancing local
31 fishery production, and protect coastal communities from SLR and storm events (IPCC, 2019b). Ecosystem-
32 based adaptation is a cost-effective coastal protection tool that can have many co-benefits, including
33 supporting livelihoods, contributing to carbon sequestration and the provision of a range of other valuable
34 ecosystem services (IPCC, 2019b).

35

36 Diversification of food systems is another component of land, ocean, and ecosystem transitions that are
37 consistent with CRD. Balanced diets, featuring plant-based foods, such as those based on coarse grains,
38 legumes, fruits and vegetables, nuts and seeds, and animal-sourced food produced in resilient, sustainable
39 and low-GHG emission manner, are major opportunities for adaptation and mitigation and improving human
40 health. By 2050, dietary changes could free several million sq. km of land and provide a mitigation potential
41 of 0.7 to 8.0 GtCO2eq yr-1, relative to business-as-usual projections.

42

43 For coastal systems, many frameworks for climate resilience and adaptation have been developed since the
44 AR5 (Hoegh-Guldberg et al., 2014; Settele et al., 2014) with substantial variations in approach between and
45 within countries, and across development status. Few studies have assessed the success of implementing
46 these frameworks due to the time-lag between implementation, monitoring, evaluation and reporting (IPCC,
47 2019g). As an example, the Nature-Based Climate Solutions for Oceans initiative has the potential to:
48 restore, protect and manage coastal and marine ecosystems, adapt to climate change, improve coastal
49 resilience, and enhance their ability to sequester and store carbon (Hoegh-Guldberg et al., 2019).

50

51 Polar regions will be profoundly different in the future. The degree and nature of that difference will depend
52 strongly on the rate and magnitude of global climate change, which will influence adaptation responses
53 regionally and worldwide. Future climate-induced changes in the polar oceans, sea ice, snow and permafrost
54 will drive habitat and biome shifts, with associated changes in the ranges and abundance of ecologically
55 important species (IPCC, 2019g). Innovative tools and practices in polar resource management and planning
56 show strong potential in improving society's capacity to respond to climate change. Networks of protected
57 areas, participatory scenario analysis, decision support systems, community-based ecological monitoring that

Do Not Cite, Quote or Distribute  18-48       Total pages: 197
    FINAL DRAFT                                  Chapter 18          IPCC WGII Sixth Assessment Report

1 draws on local and indigenous knowledge and self-assessments of community resilience contribute to
2 strategic plans for sustaining biodiversity and limit risk to human livelihoods and wellbeing. Experimenting,
3 assessing, and continually refining practices while strengthening links with decision making has the potential
4 to ready society for the expected and unexpected impacts of climate change (IPCC, 2019g).

5

6

7 [START BOX 18.6 HERE]

8

9 Box 18.6: The Role of Ecosystems in Climate-Resilient Development

10

11 Ecosystems and their services closely relate to CRD. Climate change has impacted ecosystems across a
12 range of scales, and those impacts have been exacerbated by other ecological impacts associated with human
13 activities. Ecosystem based adaptation strategies have been developed and is crucial to CRD. However,
14 knowledge and evidence still missing, and cultural services--in contrast to provision and regulation services

15 as main benefits and supporting services as co-benefits--are less well addressed in the literature.

16

17 Ecosystems play a key role in CRD

18

19 A key element of CRD is ensuring that actions taken to mitigate climate change do not compromise
20 adaptation, biodiversity, and human needs. Maintaining ecosystem health, linked to planetary health, is an
21 integral part of the goals of CRD. The 2005 Millennium Ecosystem Assessment defined ecosystem services
22 as "the benefits people obtain from ecosystems", and categorized the services in to provisioning, regulating,
23 supporting, and cultural services (Millennium Ecosystem Assessment, 2005; IPBES, 2019). The 2019
24 Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services (IPBES) broadened the
25 definition to "the contributions, both positive and negative, of living nature to the quality of life for people",
26 and developed a classification of 18 categories (IPBES, 2019).

27

28 Table Box 18.6.1 demonstrates how ecosystem services connect to sustainable development goals (SDGs)
29 and CRD. MEA's provisioning service generally connects to the IPBES' material services, mostly
30 contributing to the SDG cluster associated with nature's contribution to people (NCP) (Millennium

31 Ecosystem Assessment, 2005; IPBES, 2019) and to "Development" in CRD. MEA's regulating and

32 supporting services connect to IPBES' non-material services, contributing to SDG clusters of Nature and
33 Driver of change in nature and NCP and to "Resilience" in CRD. MEA's cultural services connect to
34 IPBES' non-material services, contributing to SDG clusters of good quality of lift (GQL) and to Enabling
35 conditions for CRD.

36

37

38 Table Box 18.6.1: Ecosystem services (based on the Millennium Ecosystem Assessment, MEA, and the

39 Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services, IPBES, classifications) and their

40 connections to sustainable development goals (SGCs) and climate resilient development (CRD) (Millennium

41 Ecosystem Assessment, 2005; IPBES, 2019).

                 Ecosystem services                                 SDGs                                  CRD
                                                                                                     Development
    MEA                  IPBES

    Provision 11 Energy                          1 No poverty

    ing  12 Food and feed                        2 Zero hunger

    services 13 Materials and assistance         3 Good health and well-being

         14 Medicinal, biochemical, and genetic 11 Sustainable cities communities

         resources                               7 Affordable clean energy

                                                 8 Decent work and economic growth

                                                 9 Industry, innovation, and infrastructure

                                                 12 Responsible consumption and

    Regulatin 3 Regulation of air quality         production                                         Climate
                                                 6 Clean water and sanitation

    g services 4 Regulation of climate           13 Climate action                                   adaptation

         5 Regulation of ocean acidification                                                         and

         6 Regulation of freshwater quantity,                                                        mitigation

         location, and timing

         7 Regulation of freshwater and coastal

         water quality

    Do Not Cite, Quote or Distribute             18-49                                       Total pages: 197
    FINAL DRAFT                                       Chapter 18                    IPCC WGII Sixth Assessment Report

              9 Regulation of hazards and extreme
               events

              10 Regulation of organisms detrimental

                   to humans                                  14 Life below water
    Supportin 1 Habitat creation and maintenance

    g services 2 Pollination and dispersal of seeds           15 Life on land

              8 Formation, protection, and

                      decontamination of soils and sediments

              18 Maintenance of options

    Cultural  15 Learning and inspiration                     4 Quality education                         Enabling

    services 16 Physical and psychological                    5 Gender equality                           Conditions

                      experiences                             10 Reduce inequality

              17 Supporting identities                        16 Peace, justice, and strong institutions

                                                              17 Partnerships for the goals

1

2

3 Climate change impacts on ecosystems and their services

 4

 5 Climate change connects to ecosystem services through two links: climate change and its influence on
 6 ecosystems as well as its influence on services (Chapter 2.2). The key climatic drivers are changes in
 7 temperature, precipitation, and extreme events, which are unprecedented over millennia and highly variable
 8 by regions (Chapter 2.3, 3.2; Cross-Chapter Box EXTREMES in Chapter 2). These climatic drivers
 9 influence physical and chemical conditions of the environment, and worsen the impacts of non-climate
10 anthropogenic drivers including eutrophication, hypoxia, sedimentation (Chapter 3.4). Such changes have
11 led to changes in terrestrial, freshwater, oceanic and coastal ecosystems at all different levels, from species
12 shifts and extinctions, to biome migration, and to ecosystem structure and processes changes (Chapter 2.4,
13 2.5, 3.4, Cross-Chapter Box MOVING PLATE in Chapter 5). Changes in ecosystems leads to changes in
14 ecosystem services including food and limber prevision, air and water quality regulation, biodiversity and
15 habitat conservation, and cultural and mental support (Chapter 2.4, 3.5). Table Box 18.6.2 presents examples
16 of climate change's impact on ecosystems and their services from other chapters in the WGII report. The
17 degradation of ecosystem services is felt disproportionately by people who are already vulnerable due to
18 historical and systemic injustices, including women and children in low-income households, Indigenous or
19 other minority groups, small-scale producers and fishing communities, and low-income countries (Chapter
20 3.5, 4.3, 5.13).

21

22

23 Table Box 18.6.2: Examples of key risks to ecosystems from climate change and their connections to ecosystem

24 services (ES) in the WGII report and cross-chapter papers (CCPs). (See Table 1 for the description of the categories of

25 ES)

              Climate factors                                        Key risk                                 ES
                                                                                                          PRS C

    Terrestrial and freshwater ecosystems (Chapter 2, 4, 5; CCP 1; CCP 7; CCP 3; CCP 5)

    - Increase in average and extreme       Species extinction and range shifts                           X XX
     temperatures                           Ecosystem structure and process change                        X X
    - Changes in precipitation amount and

     timing                                 Ecosystem carbon loss                                         X X
    - Increase in aridity

    - Increase in frequency and severity of Wildfire                                                         X X
                                                                                                          X X
     drought                                Water cycle & scarcity
    - Increased atmospheric CO2

    Ocean and coastal (Chapter 3; CCP 1; CCP 6)

    - Ocean warming                         Species extinction and range shifts                           X XX
    - Marine heatwaves                      Ecosystem structure and process change                        X X
    - Ocean acidification                   Habitat loss                                                  X X
    - Loss of oxygen

    - Sea level rise                        Ocean carbon sink less effective                                 X

    - Increased atmospheric CO2             Erosion and land loss

    - Extreme events                                                                                      X X

    Food, Fiber, and other Ecosystem Products (Chapter 5)

    - Global warming                        Species distribution                                          X

    Do Not Cite, Quote or Distribute                          18-50                                       Total pages: 197
    FINAL DRAFT                                Chapter 18               IPCC WGII Sixth Assessment Report

    - Water stress                             Timing of key biological events change                       X

    - Extreme events                           Corp productivity and quality decrease                       X

    - Ocean acidification                      Diseases and insect                                          X

    - Salt intrusion

1

2 Adaptation practices and enabling conditions for CRD

3

4 Ecosystem protection and restoration, ecosystem-based adaptation (EbA), and nature-based solution (NbS)
5 can lower climate risk to people and achieve multiple benefits including food and material provision, climate

6 mitigation, and social benefits (Chapter 2.6, 3.6, 4.6, 5.13,6.3, 8.6). Table Box 18.6.3 presents some

7 examples of ecosystem adaptation practices reported in WGII sectoral and regional chapters and CCPs, as

 8 well as their co-benefits, potential for maladaptation, and enabling conditions. Many of the strategies focus
 9 on integrated systems (managing for multiple objectives and trade-offs) as well as the fair use of resources.
10 However, there is limited evidence of the extent to which adaptation is taking place and virtually no

11 evaluation of the effectiveness of adaptation in the scientific literature (Chapter 2.6, 3.5). Enabling
12 conditions for the successful implementation ecosystem-based practice include regional and community-
13 based based approaches, multistakeholder and multi-level governance approaches, Integration of Local
14 Knowledge and Indigenous Knowledge, finance, and social equity (Chapter 2.6, 3.6).

15

16

17 Table Box 18.6.3: Examples of adaptation practices and their connections to ecosystem services (ES) and climate

18 resilient development pathways (CRDP) in the WGII sectoral and regional chapters and cross-chapter papers (CCPs).

19 (See Table 1 for the description of the categories of ES and CRDP)

        Adaptation practices (and -            Main benefit (and & co-benefit; - trade off; + enabling         ES

                      examples)                conditions; X barrier and potential maladaptation) P R S C

    Agroforestry (Table 2.7; Table 5.ES; Food provision
    Chapter 5.10.4; Chapter 5.12.5.2; Box & Fuel (wood) provision, carbon sequestration,

    5.10; Table 16.2)                          biodiversity and ecosystem conservation, diversification

        - Climate Adaptation and               and improved economic incomes, water and soil

        Maladaptation in Cocoa and Coffee conservation, and aesthetics

        Production (Box 5.7)                   + Secure tenure arrangements, supporting Indigenous *** **           **

                                               knowledge, inclusive networks and socio-cultural

                                                 values, access to information and management skill
                                               X Higher water demand; disruption of hydrology; loss of

                                               native biodiversity; reduced resilience of certain plants;

                                               degraded soil and water quality; improper and increased

                                               use of agrochemicals, pesticides, and fertilizers

    Forest maintenance and restoration         Ecosystem conservation
    (Box 2.2; Table 16.2; Table Cross-         & Food provision, fuel provision, job creation, carbon

    Chapter Box NATURAL.1 in Chapter sequestration, biodiversity conservation, air quality

    2)                                         regulation, water and soil conservation, vector-borne

        - Protected area planning in Thailand disease control, improved mental health, cultural

         (Chapter 2.6.5.3)                     benefits, natural resources relative conflict prevention
        - Conserving Joshua trees in the
         Joshua National Park (Chapter         + Cooperation of indigenous peoples and other local

         2.6.5.6)                                communities                                                ** ** *** **
        - Addressing Vulnerability of Peat     X Planting large scale non-native monocultures leads to
         Swamp Forests in South East Asia
                                               loss of biodiversity and poor climate change resilience,

                                               increased vulnerability to landslide, increased sensitivity

         (Chapter 2.6.5.10)                    of new tree species, reduced resilience of certain plants,
        - Reduce emissions from                high water demand, trees planted damaged buildings
         deforestation and forest degradation  during heavy storms, lack of carbon rights in national
         (REDD+) (Chapter 5.6.3.3; Table       legislations

        16.2)

    Traditional practices/indigenous           Food and material provision                                  *** **
    knowledge and local knowledge              & Carbon sequestration
    (IKLK) (Table 2.7; Chapter 5.6.3;          + Partnerships between key stakeholders such as
    Chapter 5.14.2.2; Table 16.2)
                                                 researchers, forest managers, and local actors,
      - Crop and livestock farmers on            indigenous and local knowledge
        observed changes in climate in the

        Sahel (Box 5.6)

    Do Not Cite, Quote or Distribute           18-51                                                        Total pages: 197
   FINAL DRAFT                                   Chapter 18               IPCC WGII Sixth Assessment Report

   - Karuk Tribe in northern California
    (Chapter 5.6.3.2)

   Restoring natural fire regimes (Table Fire regulation

   2.7)                                    & Biodiversity conservation

   - Protecting Gondwanan wildfire                                                                           ***

   refugia in Tasmania, Australia

   (Chapter 2.6.5.8)

   Natural flood risk management (Table    Water security, flood regulation, sediment retention
   2.7)                                    & Biodiversity and ecosystem conservation

     - Natural Flood Management (NFM)                                                                        *** **
       in England, United Kingdom
       (Chapter 2.6.5.2)                   Coastal protection against sea level rise and storm surges
   Coastal ecosystem conservation (Table
   Cross-Chapter Box NATURAL.1 in          & Fisheries, carbon sequestration, biodiversity and
   Chapter 2) (Table 16.2)(Table 2.7)
     - African penguin on-site adaptation  ecosystem conservation, flood regulation, water
       (Chapter 2.6.5.5)
                                             purification, recreation, and cultural benefits                 ** *** **
   Eco-tourism within protected areas      X NH4 emissions, digging channels and sand walls around
   (Table 2.7)
                                           homes, loss of recreational value of beaches, shifted the

                                           flood impacts to poor informal urban settlers, erosion

                                           and degraded coastal lands

                                           Tourism                                                      *** **
                                           & Habitat protection

   Aquaculture (Chapter 5.9.4; Table 16.2; Food provision
   Table Cross-Chapter Box NATURAL.1 & Biodiversity conservation

   in Chapter 2)                           + Farmer incentives, participatory adaptation to context
                                           X Lack of financial, technical or institutional capacity; *** *

                                           short value chains; productivity varies by system; over-

                                           fertilizing; deforestation of mangroves; salt intrusion;

                                           increased flood vulnerability

   Water-energy-food (WEF) nexus (Box Water, energy, and food provision

   4.7)                                    X Insufficient data, information, and knowledge in

   - Food Water Energy Nexus in Asia       understanding the WEF inter-linkages; lack of systematic

    (Chapter 10.6.3)                       tools to address trade-offs involved in the nexus            ***
   - New Zealand's Land, Water and
    People Nexus under a changing

   climate (Box 11.7)

   Urban greening (Table 2.7; table 16.2; Urban flood management, water savings, urban heat island

   Table Cross-Chapter Box NATURAL.1 mitigation

   in Chapter 2)                           & Reduced carbon emissions, air and noise regulation,

   - Ecosystem based adaptation in         improved mental health, energy savings, recreation, and

   Durban, South Africa (Chapter           aesthetics

   2.6.5.7)                                + Meaningful partnerships, long-term financial

                                           commitments, and significant political and                        *** **
                                           administrative

                                           X Storage of large quantities of water in the home; water

                                           contamination; increased breeding sites for mosquitoes

                                           and flies; vectors and diseases; intensified cultivation of

                                           marginal lands; clearing of virgin forests for farmland;

                                           frequent weeding; increased competition for water and

                                           nutrients; reduced soil fertility, invasive species

1

2

3 [END BOX 18.6 HERE]

4

5

6 18.3.1.4 Industrial systems

 7

 8 Industrial emissions have been growing faster since 2000 compared to emissions in any other sector, driven
 9 by increased extraction and production of basic materials (Crippa et al., 2019; IEA, 2019) (very high
10 confidence). About one-third of the total emissions are contributed by the industry sector, if indirect emissions
11 from energy use are considered (Crippa et al., 2019). The COVID-19 pandemic has caused a significant

   Do Not Cite, Quote or Distribute                    18-52                                            Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1 decrease in demand for fuels, oil, coal, gas, and nuclear energy (IEA, 2020). However, there is concern that
2 the rebound in the crisis will reverse this trend (IEA, 2020). Accordingly, the literature suggests a combined
3 set of measures is beneficial for facilitation a transition of industrial systems in support of CRD. This includes
4 (i) dematerialization and decarbonization of industrial systems, (ii) establishment of supportive governance,
5 policies, and regulations, and (iii) implementation of enabling corporate strategies.

 6 Decarbonization and dematerialization strategies have been proposed as key drivers for the transition of
 7 industrial systems (Fischedick et al., 2014; Worrell et al., 2016). The former involves limiting carbon
 8 emissions from industrial processes (IEA, 2017; Hildingsson et al., 2019), while the latter involves improving
 9 material efficiency, developing circular economies, raw material demand management, environmentally
10 friendly product and process innovations, and environmentally friendly supply chain management (Worrell et
11 al., 2016; Petrides et al., 2018).

12 Recent modelling suggests that stocks of manufactured capital, including buildings, infrastructure,
13 machinery, and equipment, stabilize as countries develop and decouple from GDP (high agreement, medium
14 evidence). For instance, Bleischwitz et al. (2018) confirmed the occurrence of a saturation effect for
15 materials in four energy-intensive sectors (steel, cement, aluminum and copper) in five industrialized
16 countries (Germany, Japan, the United Kingdom, the United States and China). High growth in the supply of
17 materials may still drive global demand for new products in the coming years for developing countries that
18 are still far from saturation levels. Therefore, accelerating industrial transitions to drive the decoupling of
19 industrial emissions from economic growth and facilitate broader transformation in industrial systems can be
20 one component of CRD.

21

22 Continued transitions in the industrial sector will be contingent on technological innovation. Although
23 technologies exist to drive emissions in industrial sectors to very low or zero emissions, but they require 5 to
24 15 years of innovation, commercialization, and intensive policies to ensure uptake (舎man et al., 2017)
25 (high agreement, medium evidence). For instance, several options exist to reduce GHG emission related to
26 steel production process including increasing the share of the secondary route (Pauliuk et al., 2013),
27 hydrogen-based direct reduced iron (Vogl et al., 2018), aqueous electrolysis rout (Cavaliere, 2019), and
28 plasma process (Quader et al., 2016).

29

30 Industrial transitions are also contingent upon consumer behavior in terms of preferences for, and rates of,
31 consumption of industrial products. Sustainable consumption can play an important role in sustainable
32 production (Allwood et al., 2013; Allwood et al., 2019). This suggests feedbacks between industrial
33 production and consumption in driving industrial transitions. For example, sustainable consumption can be
34 triggered and/or enabled through sustainable production processes that provide more sustainable options to
35 consumers as well as public or private promotional campaigns that promote those options. Meanwhile,
36 demand from consumers for more sustainable options helps to drive the expansion of markets and innovation
37 among industrial producers to meet that demand.

38

39 18.3.1.5 Societal systems

40

41 This chapter contributes a fifth system transition in addition to the four which have already been introduced
42 by SR1.5: the societal systems transition. While society and people also feature in the other systems
43 transitions, the purpose of defining a fifth transition is to explicitly highlight the challenges associated with
44 changes in behavior, attitudes, values and consciousness required to achieve CRD. One caveat of considering
45 transitions in societal systems is the limit to which the nature of change is known: transitions accomplish
46 reconfigurations towards a relatively known destination. Historical and current differences between and
47 within nations translate to a multitude of equally valid but diverse priorities for development, for example
48 the understanding of development toward progress as linear has been challenged as being a Western concept
49 by scholars of colonialization (Sultana et al., 2019). Thus societal transitions are understood as being
50 intrinsically diverse for the purpose of achieving climate resilient development.

51

52 The four systems transitions identified in SR1.5 already include a component of societal change � for
53 example, attitude change is part of public acceptance that facilitates shifts in energy including changing
54 electricity to renewables (Ch 4 SR1.5 4.3.1.1) and developing nuclear power (4.3.1.3), and behavioral
55 change is a part of shifting irrigation practices to drive required land and ecosystems transitions (4.3.2.1).

Do Not Cite, Quote or Distribute  18-53       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 Extracting societal transitions also allows for a detailed examination of other societal dimensions that
 2 facilitate systems transitions, for example justice issues relating to water and energy access and distribution,
 3 and land use. Societal transition, sometimes known as `societal transformation', is an established concept in
 4 different literatures, as described below. Transformation and transition are terms often used as synonyms
 5 (H鰈scher et al., 2018) although different schools of thought understand them as sub-components of each
 6 other, eg. transition driving transformation, or transformation driving transition. For a more detailed
 7 discussion on the differences between transition and transformation represented in the literature, see Box
 8 18.1.

 9

10 Societal transitions for the purpose of this report are understood as the collection of shifts in attitudes, values,
11 consciousness and behavior required to move toward CRD. This builds on the SR1.5 (IPCC, 2018a: 599)
12 definition of societal (social) transformation: "A profound and often deliberate shift initiated by communities
13 toward sustainability, facilitated by changes in individual and collective values and behaviors, and a fairer
14 balance of political, cultural, and institutional power in society." This includes accepting IK/LK as an
15 equally valid form of knowledge as compared with Western, scientific knowledge (see Cross-Chapter Box
16 INDIG) and recognition of the role of shifting gender norms to achieve climate resilience (see Cross-Chapter
17 Box GENDER). Changes associated with societal transitions are not specific to defined systems (e.g. energy,
18 industry, land/ecosystems or urban/infrastructure). Rather, these sectoral systems are embedded within
19 broader societal systems, including e.g. political systems, economic systems, knowledge systems, cultural
20 systems (Davelaar, 2021; Turnhout et al., 2021; Visseren-Hamakers et al., 2021). Changes that happen in
21 these broader social systems can therefore prompt changes in all systems embedded within them, meaning
22 that societal transition is key to transforming across a range of sectors and topics (Leventon et al., 2021).
23 Furthermore, societal transition requires changes in individual behaviors, but also in the broader conditions
24 that shape these behaviors. These broader conditions are largely related to questions of power, in enforcing
25 dominant political economies and social-technological mindsets (Stoddard et al., 2021). This section also
26 briefly describes the various trains of research on societal transitions and transformation.

27

28 Because of the multiple sectors, interests and scales that are involved in societal transitions, understanding
29 and creating evidence on transitions requires shifting across system boundaries and finding ways to
30 transcend disciplinary silos. Relevant research includes work within the topic of transformation and
31 transitions (H鰈scher et al., 2018). Transformations literature can be split into multiple sub-concepts and
32 requires engagement with multiple schools of thought (Feola, 2015; Feola et al., 2021). Much focus within
33 transformations research is currently related to biodiversity conservation (Massarella et al., 2021), and
34 transitions work tends towards a focus in urban areas (Loorbach et al., 2017). Though there is also work in
35 both that is more broadly labelled as sustainability transformations or transitions (Luederitz et al., 2017).
36 Furthermore, there is likely to be much relevant literature that does not explicitly label itself as
37 transformations or transitions (Feola et al., 2021). For example, we could look to political science theories on
38 policy change (Leventon et al., 2021) and historical perspectives on social change. Bridging these divides
39 will require a deeper rethinking in the research community to undo power structures that marginalize diverse
40 knowledges (Caniglia et al., 2021; Lahsen and Turnhout, 2021).

41

42 There are a number of concepts proposed as pathways to creating societal transitions; usually centered
43 around the idea of working with individuals and communities to change their mindsets as a way to change
44 the way they manage their local environments or behave. Transformations work explores how values are
45 pathways towards sustainability, for example by changing values, through making values explicit, through
46 negotiation, and by eliciting values (Horcea-Milcu et al., 2019). Human nature connections is a further
47 concept that is identified as a way to shift values and behaviors across a range of disciplines (Ives et al.,
48 2017). The role of learning and indigenous knowledge is also explored (Lam et al., 2020). These three
49 concepts have had particular salience in discussions around transformations for biodiversity conservation
50 and restoration, related to the IPBES assessment on Values (Pascual et al., 2017; Peterson et al., 2018). They
51 largely focus on the need to engage with people's values, connections and knowledge to better manage the
52 social-ecological system they are in.

53

54 Focusing on bottom-up and community-led transformations, there is emphasis on the role of grassroots
55 organizations in transformations. Community actions around specific locations or topics have parallels to the
56 idea of transformative spaces. They are sites of innovative activity (Seyfang and Smith, 2007). Grassroots
57 organizations can bridge the local and the political scales by politicizing actors and creating new interactions

Do Not Cite, Quote or Distribute  18-54       Total pages: 197
    FINAL DRAFT                       Chapter 18                     IPCC WGII Sixth Assessment Report

1 between individuals and political processes (Nov醟, 2021). They are a collective approach to pushing for
2 both individual and societal change (Sage et al., 2021).

 3

 4 Despite a current lack of empirical evidence, there are numerous frameworks emerging for exploring societal
 5 transitions across levels. There is focus on pathways for sustainability transitions, which tends to look at
 6 projected, normative scenarios for the future, and explore or back-cast the institutional and societal changes
 7 that are required to get there (Westley et al., 2011; Sharpe et al., 2016). There is also work that looks at
 8 scaling up of smaller sustainability initiatives, through processes of scaling up, scaling out and scaling deep
 9 (Moore et al., 2015; Lam et al., 2020). In particular, systems thinking provides an organizing framework for
10 bringing together multiple disciplines and perspectives, to understand problem framings, and normative and
11 design aspects of social systems and behaviors (Foster-Fishman et al., 2007). Within this, Meadows (1999)
12 framework of leverage points for systems transformation has been operationalized within the sustainability
13 transformations debate (Abson et al., 2017). Here, system properties relating to system paradigms and design
14 are leverage points where interventions can create greatest system change; shallower leverage points relate to
15 materials and processes. This framework is increasingly being used across a range of sustainability problems
16 as boundary objects for cross-disciplinary, critical research (Fischer and Riechers, 2019; Leventon et al.,
17 2021; Riechers et al., 2021).

18

19 Analyses of societal transitions have had limited engagement with adaptation questions. The focus of the
20 sub-field of sustainability transitions on a few industrialized nations, mostly in North America and Europe,
21 limited the field's development to assumptions born from the experiences in those areas. More recent studies
22 have sought to understand sustainability transitions in other countries, especially emerging economies
23 (Wieczorek, 2018; K鰄ler et al., 2019). In particular, China has received attention from scholars on
24 sustainability transitions (Huang et al., 2018; Lo and Cast醤 Broto, 2019; Cast醤 Broto et al., 2020; Huang
25 and Sun, 2020). As a result, some pressing issues related to societal transitions for adaptation have received
26 limited attention compared with that paid to other system transitions. However, more recently, scholarship
27 has begun examining transitions that have turned to nature and nature-based solutions. Adaptive transitions
28 are an intermediary step towards sustainability transitions whereby multiple actions at material and
29 institutional levels are combined towards improving adaptation outcomes (Pant et al., 2015; Scarano, 2017).

30

31

32 Table 18.3: Specific options for facilitating the five system transitions that can support CRD

      Transition                           Examples                                Reference
    Energy                                                           (Gielen et al., 2019) (Mulugetta et
    Systems         � Fuel switching from coal to natural gas        al., 2019) (IEA et al., 2019) AR6
                    � Expansion of renewable energy technologies     WGIII Chapter 2
                    � Financial incentives to promote renewable

                        energy
                    � Reduced energy intensity of industry
                    � Improvements in power system resilience

                        and reliability
                    � Increased water use efficiency in electricity

                    generation

                    � Energy demand management strategies

    Urban and       � Increased investment in physical and social (IPCC, 2018b): D3.1)
    infrastructure
    systems         infrastructure

    Land, Oceans,   � Enhance urban and regional planning
    and             � Enhanced governance and institutional
    Ecosystems
                    capacity supports post-disaster recovery and

                    reconstruction (Kull, 2016)

                    � Expanding access to agricultural and climate (IPCC, 2019f): C2.1) (IPCC,

                    services                                         2019f): C4.5) (IPCC, 2019f): C4)

                    � Strengthening land tenure security and access

                    to land

                    � Empowering women farmers
                    � Improved access to markets
                    � Facilitating payments for ecosystem services

    Do Not Cite, Quote or Distribute             18-55                                             Total pages: 197
FINAL DRAFT                              Chapter 18           IPCC WGII Sixth Assessment Report

 Industrial  � Promotion of healthy and sustainable diets
 systems
             � Enhancing multi-level governance by

             supporting local management of natural

             resources

             � Strengthening cooperation between

             institutions and actors

             � Building on local, indigenous and scientific

             knowledge funding, and institutional support

             � Monitoring and forecasting

             � Education and climate literacy and social

             learning and participation

             � Promote material efficiency and high-quality (舎man et al., 2017; Bataille et al.,

             circularity                                      2018; Material, 2019) (Tanaka,

             � Materials demand management (IEA 2019, 2011; Schwarz et al., 2020)

             2020)                                            (Ciwmb, 2003) (Romero

             � Application of new processes and               Mosquera, 2019) (Tanaka, 2011)

             technologies for GHG emission reduction (Ryan et al., 2011; Boyce, 2018)
                                                              (Taylor, 2008) (UNEP, 2018b)
             � Carbon pricing or regulations with
                                                              (Kaza et al., 2018) (S鰀erholm and
             provisions on competitiveness to drive
             innovation and systemic carbon efficiency Tilton, 2012) (Bataille et al., 2018)
             � Low-cost, long-term financing mechanisms (Ghisetti et al., 2017) (Taylor,
                                                              2008; Fischedick et al., 2014;
             to enable investment and reduce risk
                                                              Hansen and Lema, 2019) (Crippa
             � Better planning of transport infrastructure    et al., 2019; IEA, 2019) (Cavaliere,
             � Labour market training and transition          2019; IEA, 2020)(Vogl et al.,
                                                              2018)(Pauliuk et al., 2013; Quader
                 support                                      et al., 2016)
             � Electricity market reform
             � Regulations � standards and labelling,

             material efficiency

             � Mandating technologies and targets

             � Green taxes and carbon pricing, preferential

             loans and subsidies

             � voluntary action agreements, expanded

             producer responsibilities

             � information programs: monitoring,

             evaluation, partnerships, and research and

             development

             � government provisioning of services--

             government procurements, technology push

             and market-pull

Societal     � Inclusive governance                           (Fazey et al., 2018b; O'Brien,
Systems      � Empowerment of excluded stakeholders,          2018; Patterson et al., 2018)
                                                              (MRFCJ, 2015; Dumont et al.,
                 especially women and youth                   2019) (Popescu et al., 2017;
             � transforming economies                         David T郻ara et al., 2018) (de
             � finance and technology aligned with local      Coninck and Sagar, 2015; IEA,
                                                              2015; Parikh et al., 2018)
                 needs                                        (Dearing et al., 2014; H鋣h� et al.,
             � overcoming uneven consumption and              2016; Raworth, 2017) (Klinsky
                                                              and Winkler, 2018), (Hajer et al.,
                 production patterns                          2015; Labriet et al., 2015; Hale,
             � allowing people to live a life in dignity and  2016; Pelling et al., 2016;
                                                              Kalafatis, 2017; Lyon, 2018)
                 enhancing their capabilities                 (Holden et al., 2017) (Cundill et
             � involving local governments, enterprises       al., 2014; Butler et al., 2016;
                                                              Ensor, 2016; Fazey et al., 2016;
                 and civil society organisations across

                 different scales

Do Not Cite, Quote or Distribute           18-56              Total pages: 197
    FINAL DRAFT                        Chapter 18                IPCC WGII Sixth Assessment Report

                 � reconceptualising development around          Gorddard et al., 2016; Aipira et
                                                                 al., 2017; Chung Tiam Fook,
                     well-being rather than economic growth      2017; Maor et al., 2017) (O'Brien
                                                                 and Selboe, 2015; Gillard et al.,
                     (Gupta and Pouw, 2017),                     2016; DeCaro et al., 2017; Harris
                 � rethinking, prevailing values, ethics and     et al., 2018; Lahn, 2018; Roy et
                                                                 al., 2018) Sections 5.6.1 and
                     behaviour                                   5.5.3.1
                 � improving decision-making processes that

                     incorporate diverse values and world views
                 � creating space for negotiating diverse

                     interests and preferences

1

2

3 [START CROSS-CHAPTER BOX GENDER HERE]

4

5 Cross-Chapter Box GENDER: Gender, Climate Justice and Transformative Pathways

 6

 7 Authors: Anjal Prakash (India), Cecilia Conde (Mexico), Ayansina Ayanlade (Nigeria), Rachel Bezner Kerr
 8 (Canada/USA), Emily Boyd (Sweden), Martina A Caretta (Sweden), Susan Clayton (USA), Marta G. Rivera
 9 Ferre (Spain), Laura Ramajo Gallardo (Chile), Sharina Abdul Halim (Malaysia), Nina Lansbury (Australia),
10 Oksana Lipka (Russia), Ruth Morgan (Australia), Joyashree Roy (India), Diana Reckien
11 (Netherlands/Germany), E. Lisa F. Schipper (Sweden/UK), Chandni Singh (India), Maria Cristina Tirado
12 von der Pahlen (Spain/USA), Edmond Totin (Benin), Kripa Vasant (India), Morgan Wairiu (Solomon
13 Islands), Zelina Zaiton Ibrahim (Malaysia).

14

15 Contributing Authors: Seema Arora-Jonsson (Sweden/India), Emily Baker (USA), Graeme Dean (Ireland),
16 Emily Hillenbrand (USA), Alison Irvine (Canada), Farjana Islam (Bangladesh/ United Kingdom), Katriona
17 McGlade (United Kingdom/Germany), Hanson Nyantakyi-Frimpong (Ghana), Nitya Rao (United Kingdom
18 /India), Federica Ravera (Italy), Emilia Reyes (Mexico), Diana Hinge Salili (Fiji), Corinne Schuster-Wallace
19 (Canada), Alcade C. Segnon (Benin), Divya Solomon (India), Shreya Some (India), Indrakshi Tandon
20 (India), Sumit Vij (India), Katharine Vincent (United Kingdom/South Africa), Margreet Zwarteveen (The
21 Netherlands)

22

23 Key Messages

24

25  Gender and other social inequities (e.g., racial, ethnic, age, income, geographic location) compound

26  vulnerability to climate change impacts (high confidence). Climate justice initiatives explicitly address

27  these multi-dimensional inequalities as part of a climate change adaptation strategy. [Box 9.2:

28  Vulnerability Synthesis: Differential Vulnerability by Gender and Age in Ch 9]

29

30  Addressing inequities in access to resources, assets, and services as well as participation in decision-

31  making and leadership is essential to achieving gender and climate justice (high confidence).

32

33  Intentional long-term policy and program measures and investments to support shifts in social rules,

34  norms, and behaviours are essential to address structural inequalities and support an enabling

35  environment for marginalised groups to effectively adapt to climate change (very high confidence).

36  [Equity and Justice box in Ch 17]

37

38  Climate adaptation actions are grounded in local realities so understanding links with SDG 5 is

39  important to ensure that adaptive actions do not worsen existing gender and other inequities within

40  society (e.g., leading to maladaptation practices) (high confidence). [17.5.1]

41

42  Adaptation actions do not automatically have positive outcomes for gender equality. Understanding the

43  positive and negative links of adaptation actions with gender equality goals, (i.e., SDG 5), is important to

44  ensure that adaptive actions do not exacerbate existing gender-based and other social inequalities

45  [16.1.4.4]. Efforts are needed to change unequal power dynamics and 磘o foster inclusive decision-

46  making for climate adaptation to have a positive impact for gender equality (high confidence).

47

48  There are very few examples of successful integration of gender and other social inequities in climate

49  policies to address climate change vulnerabilities and questions of social justice, (Very high confidence).

    Do Not Cite, Quote or Distribute   18-57                                        Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1
 2

 3 Gender, climate justice, and climate change

 4

 5 This Cross-Chapter Box highlights the intersecting issues of gender, climate change adaptation, climate
 6 justice, and transformative pathways. A gender perspective does not centre only on women or men but
 7 examines structures, processes, and relationships of power between and among groups of men and women
 8 and how gender, particularly in its non-binary form, intersects with other social categories such as race,
 9 class, socio-economic status, nationality, or education to create multidimensional inequalities (Hopkins,
10 2019). A gender transformative approach aims to change structural inequalities. Attention to gender in
11 climate change adaptation is thus central to questions of climate justice that aim for a radically different
12 future (Bhavnani et al., 2019). As a normative concept highlighting the unequal distribution of climate
13 change impacts and opportunities for adaptation and mitigation, climate justice (Wood, 2017; Jafry et al.,
14 2018; Chu and Michael, 2019; Shi, 2020a) calls for transformative pathways for human and ecological
15 wellbeing. These address the concentration of wealth, unsustainable extraction, and distribution of resources
16 (Schipper et al., 2020a; Vander Stichele, 2020) as well as the importance of equitable participation in
17 environmental decision-making for climate justice (Arora-Jonsson, 2019).

18

19 Research on gender and climate change demonstrates that an understanding of gendered relations is central
20 to addressing the issue of climate change. This is because gender relations mediate experiences with climate
21 change, whether in relation to water (K鰄ler et al., 2019) (see also Sections 4.7, 4.3.3; 4.6.4, 5.3), forests
22 (Arora-Jonsson, 2019), agriculture (Carr and Thompson, 2014; Balehey et al., 2018; Garcia et al., 2020) (see
23 also Chapter 4, Section 5.4), marine systems (Mcleod et al., 2018; Garcia et al., 2020) (see also Section 5.9)
24 or urban environments (Reckien et al., 2018; Susan Solomon et al., 2021) (see also Chapter 6). Climate
25 change has direct negative impacts on women's livelihoods due to their unequal control over and access to
26 resources (e.g., land, credit) and because they are often the ones with the least formal protection (Eastin,
27 2018) (see also Box 9.2 in Ch 9). Women represent 43% of the agricultural labour force globally, but only
28 15% of agricultural landholders (OECD, 2019b). Gendered and other social inequities also exist with non-
29 land assets and financial services (OECD, 2019b) often due to social norms, local institutions, and
30 inadequate social protection (Collins et al., 2019b). Men may experience different adverse impacts due to
31 gender roles and expectations (Bryant and Garnham, 2015; Gonda, 2017). These impacts can lead to
32 irreversible losses and damages from climate change across vulnerability hotspots (Section 8.3).

33

34 Participation in environmental decision-making tends to favour certain social groups of men, whether in
35 local environmental committees, international climate negotiations (Gay-Antaki and Liverman, 2018) or the
36 IPCC (Nhamo and Nhamo, 2018). Addressing climate justice reinforces the importance of considering the
37 legacy of colonialism on developing regional and local adaptation strategies. Scholars have criticized climate
38 programs for setting aside forestland that poor people rely on and appropriating the labor of women in the
39 global South without compensatory social policy or rights; where women are expected to work with Non
40 Timber Forest Products to compensate for the lack of logging and for global climate goals but where their
41 work of social reproduction and care is paid little attention (Westholm and Arora-Jonsson, 2015; Arora-
42 Jonsson et al., 2016). A global ecologically unequal exchange, biopiracy, damage from toxic exports, or the
43 disproportionate use of carbon sinks and reservoirs by high-income countries enhance the negative impacts
44 of climate change, women in LDC's and SIDS also endure the harshest impacts of the debt crisis due to
45 imposed debt measures in their countries (Appiah and Gbeddy, 2018; Fresnillo Sallan, 2020). The austerity
46 measures derived as conditionalities for fiscal consolidation in public services increases gender-based
47 violence (Casta馿da Carney et al., 2020) and brings additional burdens for women in the form of increasing
48 unpaid care and domestic work (Bohoslavsky, 2019).

49

50 Gendered vulnerability

51

52 Land, ecosystem, and urban transitions to climate-resilient development need to address gender and other
53 social inequities to meet sustainability and equity goals, otherwise, marginalised groups may continue to be
54 excluded from climate change adaptation. In the water sector, increasing floods and droughts and
55 diminishing groundwater and runoff have gendered effects on both production systems and domestic use
56 (Sections 4.3.1, 4.3.3, 4.5.3). Climate change is reducing the quantity and quality of safe water available in
57 many regions of the world and increasing domestic water management responsibilities (high confidence). In
58 regions with poor drinking water infrastructure, it is forcing, primarily women and girls, to walk long

Do Not Cite, Quote or Distribute  18-58       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 distances to access water, and limiting time available for other activities, including education and income
 2 generation (Eakin et al., 2014; Kookana et al., 2016; Yadav and Lal, 2018). Water insecurity and the lack of
 3 water, sanitation, and hygiene (WASH) infrastructure have resulted in psychosocial distress, gender-based
 4 violence, as well as poor maternal and child health and nutrition (Collins et al., 2019a; Wilson et al., 2019;
 5 Geere and Hunter, 2020; Islam et al., 2020; Mainali et al., 2020) (Sections 4.3.3 and 4.6.4.4) (high
 6 confidence). Climate-related extreme events also affect women's health � by increasing the risk of maternal
 7 and infant mortality, disrupting access to family planning and prevention of mother to child transmission
 8 regimens for HIV positive pregnant women (Undrr, 2019) (see also Section 7.2). Women and the elderly are
 9 also disproportionately affected by heat events (Section 7.1.7.2.1, 7.1.7.2.3, 13.7.1).

10

11 Extreme events impact food prices and reduce food availability and quality, especially affecting vulnerable
12 groups, including low-income urban consumers, wage labourers, and low-income rural households who are
13 net food buyers (Green et al., 2013; Fao, 2016) (Section 5.12). Low-income women, ethnic minorities, and
14 Indigenous communities are often more vulnerable to food insecurity and malnutrition from climate change
15 impacts, as poverty, discrimination, and marginalisation intersect in their cases (Vinyeta et al., 2016; Clay et
16 al., 2018) (Section 5.12). Increased domestic responsibilities of women and youth, due to migration of men,
17 can increase their vulnerability due to their reduced capacity for investment in off-farm activities and
18 reduced access to information (Sugden et al., 2014; O'Neil et al., 2017) (Section 4.3; 4.6) (high confidence).

19

20 In the forest sector, the increased frequency and severity of drought, fires, pests and diseases, and changes to
21 growing seasons, has led to reduced harvest revenues, fluctuations in timber supply and availability of wood
22 (Lamsal et al., 2017; Fadrique et al., 2018; Esquivel-Muelbert et al., 2019). Climate programs in the global
23 South such as REDD+ have led to greater social insecurity and the conservation of the forests have led to
24 more pressure on women to contribute to household incomes but without enough supporting market access
25 mechanisms or social policy (Westholm and Arora-Jonsson, 2015; Arora-Jonsson et al., 2016). In countries
26 in the global North, reduced harvestable wood and revenues have led to employment restructuring that has
27 important gendered effects and negatively affects community transition opportunities (Reed et al., 2014).

28

29 Integrating gender in climate policy and practice

30

31 Climate change policies and programs across regions reveal wide variation in the degree and approach to
32 addressing gender inequities (see Table SMCCB GENDER.2). In most regions where there are climate
33 change policies that consider gender, they inadequately address structural inequalities t resulting from
34 climate change impacts, or how gender and other social inequalities can compound risk (high confidence).
35 Experiences show that it is more frequent to address specific gender inequality gaps in access to resources.
36 Regionally, Central and South American countries (section 12.5.8) have a range of gender-sensitive or
37 gender-specific policies such as the intersectoral coordination initiative Gender and Climate Change Action
38 Plans (PAGcc), adopted in Per�, Cuba, Costa Rica, and Panam� (Casas Varez, 2017), or the Gender
39 Environmental policy in Guatemala that has a focus on climate change (B醨cena-Mart韓 et al., 2021).
40 However, countries often have limited commitment and capacity to evaluate the impact of such policies
41 (Tramutola, 2019). In North and South America, policies have failed to address how climate change
42 vulnerability is compounded by the intersection of race, ethnicity, and gender (Radcliffe, 2014; Vinyeta et
43 al., 2016) (see also section 14.6.3). gender is rarely discussed in African national policies or programmes
44 beyond the initial consultation stage (Holvoet and Inberg, 2014; Mersha and van Laerhoven, 2019), although
45 there are gender and climate change action strategies in countries such as Liberia, Mozambique, Tanzania,
46 and Zambia (Mozambique and IUCN, 2014; Zambia and IUCN, 2017). European climate change adaptation
47 strategies and policies are weak on gender and other social equity issues (Allwood, 2014; Boeckmann and
48 Zeeb, 2014; Allwood, 2020), while in Australasia, there is a lack of gender-responsive climate change
49 policies. In Asia, there are several countries that recognize gendered vulnerability to climate change (Jafry,
50 2016; Singh et al., 2021b), but policies tend to be gender-specific, with a focus on targeting women, for
51 example in the national action plan on climate change as in India (Roy et al., 2018) or in national climate
52 change plan as in Malaysia (Susskind et al., 2020).

53

54 Potential for Change and Solutions

55

56 The sexual division of labour, systemic racism and other social structural inequities lead to increased
57 vulnerabilities and climate change impacts for social groups such as women, youth, Indigenous peoples,

Do Not Cite, Quote or Distribute  18-59       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 ethnic minorities. Their marginal positions not only affect their lives negatively but their work in
 2 maintaining healthy environments is ignored and invisible in policy affecting their ability to work towards
 3 sustainable adaptation and aspirations in the SDGs (Arora-Jonsson, 2019). However, attention to the
 4 following has the potential to bring about change:

 5

 6 Creation of new, deliberative policy-making spaces that support inclusive decision-making processes and
 7 opportunities to (re)negotiate pervasive gender and other social inequalities in the context of climate change
 8 for transformation (Tschakert et al., 2016; Harris et al., 2018; Ziervogel, 2019; Garcia et al., 2020). (high
 9 confidence)

10

11 Increased access to reproductive health and family planning services, which contributes to climate change
12 resilience and socio-economic development through improved health and well-being of women and their
13 children, including increased access to education, gender equity, and economic status (Onarheim et al., 2016;
14 Starbird et al., 2016; Lopez-Carr, 2017; Hardee et al., 2018) (Sections 7.4) (high confidence).

15

16 Engagement with women's collectives is important for sustainable environments and better climate decision-
17 making whether at the global, national, or local levels (Westholm and Arora-Jonsson, 2018; Agarwal, 2020).
18 The work of such collectives in maintaining their societies and environments and in resisting gendered and
19 community violence is unacknowledged (Jenkins, 2017; Arora-Jonsson, 2019) but is indispensable
20 especially when combined with good leadership, community acceptance, and long-term economic
21 sustainability (Chu, 2018; Singh, 2019) (Section 4.6.4). Networking by gender experts in environmental
22 organizations and bureaucracies has also been important for ensuring questions of social justice (Arora-
23 Jonsson and Sijapati, 2018).

24

25 Investment in appropriate reliable water supplies, storage techniques, and climate-proofed WASH
26 infrastructure as key adaptation strategies that reduce both burdens and impacts on women and girls (Alam et
27 al., 2011; Woroniecki, 2019) (Sections 4.3.3 and 4.6.44).

28

29 Improved gender-sensitive early warning system design and vulnerability assessments to reduce
30 vulnerabilities, prioritising effective adaptation pathways to women and marginalized groups (Mustafa et al.,
31 2019; Tanner et al., 2019; Werners et al., 2021).

32

33 Established effective social protection, including both cash and food transfers, such as the universal public
34 distribution system (PDS) for cereals in India, or pensions and social grants in Namibia, that have been
35 demonstrated to contribute towards relieving immediate pressures on survival and support processes at the
36 community level, including climate effects (Kattumuri et al., 2017; Lindoso et al., 2018; Rao et al., 2019a;
37 Carr, 2020).

38

39 Strengthened adaptive capacity and resilience through integrated approaches to adaptation that include social
40 protection measures, disaster risk management, and ecosystem-based climate change adaptation (high
41 confidence), particularly when undertaken within a gender-transformative framework (Gumucio et al., 2018;
42 Bezner Kerr et al., 2019; Deaconu et al., 2019) (Cross-Chapter Box NATURAL in Chapter 2, Section 5.12,
43 Section 5.14).

44

45 For example, gender-transformative and nutrition-sensitive agroecological approaches strengthen adaptive
46 capacities and enable more resilient food systems by increasing leadership for women and their participation
47 in decision-making and a gender-equitable domestic work (high confidence) (Gumucio et al., 2018; Bezner
48 Kerr et al., 2019; Deaconu et al., 2019) (Cross-Chapter Box NATURAL in Chapter 2, Section 5.12, Section
49 5.14)

50

51 New initiatives such as the Sahel Adaptive Social Protection Program represent an integrated approach to
52 resilience that promotes coordination among social protection, disaster risk management, and climate change
53 adaptation. Accompanying measures including, health, education, nutrition, family planning, among others
54 (Daron et al., 2021).

55

56 Climate change adaptation and SDG 5

57

Do Not Cite, Quote or Distribute  18-60       Total pages: 197
    FINAL DRAFT                                      Chapter 18       IPCC WGII Sixth Assessment Report

 1 Adaptation actions may reinforce social inequities, including gender unless explicit efforts are made to
 2 change (Nagoda and Nightingale, 2017; Garcia et al., 2020) (high evidence and high agreement).
 3 Participation in climate action increases if is inclusive and fair (Huntjens and Zhang, 2016). Roy et al. (2018)
 4 assessed links among various SDGs and mitigation options. Adaptation actions are grounded in local
 5 realities especially in terms of their impacts so understanding links with the goals of SDG 5 becomes more
 6 important to make sure that adaptive actions do not worsen prevalent gender and other social inequities
 7 within society (high evidence, high agreement). In the IPCC 1.5癈 Special Report, Roy et al. (2018)
 8 assessed links between various SDGs and mitigation options, adaptation options were not considered. The
 9 current SDG 13 climate action targets do not specifically mention gender as a component for action, which
10 makes it even more imperative to link SDG 5 targets and other gender-related targets to adaptive actions
11 under SDG 13 to ensure that adaptation projects are synergistic rather than maladaptive (16.3.2.6, Table
12 16.6) (Susan Solomon et al., 2021).

13

14 This assessment is based on a systematic rapid review of scientific publications (McCartney et al., 2017;
15 Liem et al., 2020) published on adaptation actions in 9 sectors from 2014 to 2020 (see Table SMCCB
16 GENDER.1) and how they integrated gender perspectives impacting gender equity. The assessment is based
17 on over 17,000 titles and abstracts that were initially found through keyword search and were reviewed.
18 Finally, 319 relevant papers on case studies, regional assessments, and meta-reviews were assessed. Gender
19 impact was classified by various targets under SDG 5. Following the approach taken in Roy et al. (2018) and
20 (Hoegh-Guldberg et al., 2019), the linkages were classified into synergies (positive impacts or co-benefits)
21 and trade-offs (negative impacts) based on the evidence obtained from the literature review which is finally
22 used to develop net impact (positive or negative) scores (See Table Cross-Chapter Box GENDER.1 and
23 Supplementary Material)

24

25

26 Table Cross-Chapter Box GENDER.1: Interrelations between SDG5 (gender equality) and adaptation initiatives
27 in 9 major sectors

                                         Adaptation categories

    Sector                               Ecosystem-  Technological                   Behavioural
                                         based       /infrastructure                 / cultural
                                                     /information     Institutional

    Terrestrial & freshwater
    ecosystem

    Ocean & coastal ecosystem

    Mountain ecosystem

    Food, fibre & others

    Urban water & sanitation

    Poverty, livelihood &
    Sustainable Development

    Cities, settlement & key
    infrastructure
    Health, well-being, and
    changing communities'
    structure

    Industrial system transition

28

    Colour code  Description                                     Confidence levels  Symbol
                                                                 Very High
                 All net positive links                          High

                 All net negative links

    Do Not Cite, Quote or Distribute                 18-61                                  Total pages: 197
FINAL DRAFT                         Chapter 18                                      IPCC WGII Sixth Assessment Report

             Number of net positive links > number of net negative links  Medium

             Number of net negative links > number of net positive links  Low

             no literature/options                                        Very low

 1 Table Notes:
 2 Potential net synergies and trade-offs between a sectoral portfolio of adaptation actions and SDG 5 are shown. Colour
 3 codes showing the relative strength of net positive and net negative impacts and confidence levels. The strength of net
 4 positive and net negative connections across all adaptation actions within a sector are aggregated to show sector-
 5 specific links. The links are only one-sided on how adaptation action is linked to gender equality (SDG5) targets and
 6 not vice versa. Adaptation options assessed in Ecosystem-based actions are: 22 in number, options in Technological
 7 /infrastructure /information are 10, in Institutional are 17 and in Behavioural/ cultural are 13. The assessment presented
 8 here is based on literature presenting impacts on gender equality and equity of various adaptation actions implemented
 9 in various local contexts and in regional climate change policies (Table SMCCB GENDER.2).

10
11

12 Adaptation actions being implemented in each sector in different local contexts can have positive (synergies)
13 or negative (trade-offs) effects with SDG5. This can potentially lead to net positive or net negative
14 connections at an aggregate level. How they are finally realized depends on how they are implemented,
15 managed, and combined with various other interventions in particular, place-based circumstances.
16 Ecosystem-based adaptation actions and terrestrial & freshwater ecosystems have higher potential for net
17 positive connections (Roy et al., 2018) (Table Cross-Chapter Box GENDER.1 and Supplementary Material).
18 Adaptation in terrestrial and freshwater ecosystems has the strongest net positive links with all SDG-5
19 targets (medium evidence, low agreement). For example, community-based natural resource management
20 increases the participation of women, especially when they are organised into women's groups (Pineda-
21 L髉ez et al., 2015; de la Torre-Castro et al., 2017) (Supplementary Material). For poverty, livelihood and
22 sustainable development sector adaptation actions have generated more net negative scores (low evidence,
23 low agreement) (Table Cross-Chapter Box GENDER.1). For example, patriarchal institutions and structural
24 discriminations curtail access to services or economic resources as compared to men, including less control
25 over income, fewer productive assets, lack of property rights, as well as less access to credit, irrigation,
26 climate information, and seeds which devaluate women's farm-related adaptation options (Adzawla et al.,
27 2019; Friedman et al., 2019; Ullah et al., 2019) (Supplementary Material).

28

29 Among the adaptation actions, ecosystem-based actions have the strongest net positive links with SDG-5
30 targets (Table Cross-Chapter Box GENDER.1, Table SMCCB GENDER.1). In the health, well-being and
31 changing communities' sector, this is with high evidence and medium agreement, while in all other sectors
32 there is medium evidence and low agreement. Net negative links are most prominent in institutional
33 adaptation actions (Table Cross-Chapter Box GENDER.1). For example, in mountain ecosystems, changes
34 in gender roles in response to climatic and socioeconomic stressors is not supported by institutional
35 practices, mechanisms, and policies that remain patriarchal (Goodrich et al., 2019). Additionally, women
36 often have less access to credit for climate change adaptation practices, including post-disaster relief, for
37 example, to deal with salinization of water or flooding impacts (Hossain and Zaman 2018). Lack of
38 coordination among different city authorities can also limit women's contribution in informal settlements
39 towards adaptation. Women are typically underrepresented in decision-making on home construction and
40 planning and home-design decisions in informal settlements, but examples from Bangladesh show they play
41 a significant role in adopting climate-resilient measures (e.g., the use of corrugated metal roofs and partitions
42 which is important in protection from heat) (Jabeen, 2014; Jabeen and Guy, 2015; Araos et al., 2017; Susan
43 Solomon et al., 2021).

44

45 Towards climate-resilient, gender-responsive transformative pathways

46

47 The climate change adaptation and gender literature call for research and adaptation interventions that are
48 'gender-sensitive' (Jost et al., 2016; Thompson-Hall et al., 2016; Kristjanson et al., 2017; Pearce et al.,
49 2018a) and "gender-responsive", as established in Article 7 of the Paris Agreement (UNFCCC, 2015). In
50 addition, attention is drawn to the importance of `mainstreaming' gender in climate/development policy
51 (Alston, 2014; Rochette, 2016; Mcleod et al., 2018; Westholm and Arora-Jonsson, 2018). Many calls have
52 been made to consider gender in policy and practice (Ford et al., 2015; Jost et al., 2016; Rochette, 2016;
53 Thompson-Hall et al., 2016; Kristjanson et al., 2017; Mcleod et al., 2018; Lau et al., 2021; Singh et al.,
54 2021b). Rather than merely emphasising the inclusion of women in patriarchal systems, transforming

Do Not Cite, Quote or Distribute    18-62                                           Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 systems that perpetuate inequality can help to address broader structural inequalities not only in relation to
 2 gender but also other dimensions such as race and ethnicity (Djoudi et al., 2016; Pearse, 2017; Gay-Antaki,
 3 2020). Adaptation researchers and practitioners play a critical role here and can enable gender-
 4 transformative processes by creating new, deliberative spaces that foster inclusive decision-making and
 5 opportunities for renegotiating inequitable power relations (Tschakert et al., 2016; Ziervogel, 2019; Garcia et
 6 al., 2020).

 7

 8 To date, empirical evidence on such transformational change is sparse, although there is some evidence of
 9 incremental change (e.g., increasing women's participation in specific adaptation projects, mainstreaming
10 gender in national climate policies). Even when national policies attempt to be more gendered, there is
11 criticism that they use gender-neutral language or include gender analysis without proposing how to alter
12 differential vulnerability (Mersha and van Laerhoven, 2019; Singh et al., 2021b). More importantly, the mere
13 inclusion of women and men in planning does not necessarily translate to substantial gender-transformative
14 action, for example in National Adaptation Programmes of Action across sub-Saharan Africa (Holvoet and
15 Inberg, 2014; Nyasimi et al., 2018) and national and sub-national climate action plans in India (Singh et al.,
16 2021b). Importantly, there is often an overemphasis on the gender binary (and household headship as an
17 entry point), which masks complex ways in which marginalisation and oppression can be augmented due to
18 the interaction of gender with other social factors and intra-household dynamics (Djoudi et al., 2016;
19 Thompson-Hall et al., 2016; Rao et al., 2019a; Lau et al., 2021; Singh et al., 2021b).

20

21 Climate justice and gender transformative adaptation can provide multiple beneficial impacts that align with
22 sustainable development. Addressing poverty (SDG 1), energy poverty (SDG 7), WaSH (SDG 6), health
23 (SDG 3), education (SDG 4) and hunger (SDG 2) along with inequalities (SDG 5 and SDG 10) - improves
24 resilience to climate impacts for those groups that are disproportionately affected (women, low-income and
25 marginalised groups). Inclusive and fair decision-making can enhance resilience (SDG 16; Section 13.4.4),
26 although adaptation measures may also lead to resource conflicts (SDG 16; Section 13.7). Nature-based
27 solutions attentive to gender equity also support ecosystem health (SDGs 14 and 15) (Dzebo et al., 2019).
28 Gender and climate justice will be achieved when the root causes of global and structural issues are
29 addressed, challenging unethical and unacceptable use of power for the benefit of the powerful and elites
30 (MacGregor, 2014; Wijsman and Feagan, 2019; Vander Stichele, 2020). Justice and equality need to be at
31 the centre of climate adaptation decision-making processes. A transformative pathway needs to include the
32 voice of the disenfranchised (MacGregor, 2020; Schipper et al., 2020a).

33

34 [END CROSS-CHAPTER BOX GENDER HERE]

35

36

37 18.3.2 Accelerating Transitions

38

39 Successfully implementing climate actions and managing trade-offs between mitigation, adaptation and
40 sustainable development (18.2.4) has important time considerations that imply significant urgency, making
41 substantive progress in system transitions critical for CRD. Both the SDGs and the Sendai Framework, for
42 example, have target dates of 2030. Meanwhile, the Paris Agreement sets specific time horizons for NDCs
43 and the SR1.5 indicated that limiting warming to 1.5癈 would similarly require substantial climate action by
44 2030 (IPCC, 2018a). While the literature is unambiguous regarding the need for significant system
45 transitions to achieve CRD (Section 18.1.3), the current pace of global emissions reductions, poverty
46 alleviation, and development of equitable systems of governance is incommensurate with these policy time
47 tables (Rogelj et al., 2010; Burke et al., 2016; Oleribe and Taylor-Robinson, 2016; Kriegler et al., 2018;
48 Frank et al., 2019; Sadoff et al., 2020). As noted previously in the AR5, "delaying action in the present may
49 reduce options for climate-resilient pathways in the future" (Denton et al., 2014: 1123). Accordingly,
50 significant acceleration in the pace of system transitions is necessary to enable the implementation of
51 mitigation, adaptation, and sustainable development initiatives consistent with CRD (very high confidence).

52

53 Studies since the AR5 directly address the issue of how to accelerate transitions within the broader system
54 transitions, sustainability transitions, and socio-technical transitions literature (Frantzeskaki et al., 2017;
55 Gliedt et al., 2018; Gorissen et al., 2018; Johnstone and Newell, 2018; Kuokkanen et al., 2019; Markard et
56 al., 2020). Such literature explores several core themes to facilitate acceleration, which are aligned with the
57 discussion later in this chapter on arenas of engagement for CRD (Section 18.4.3). One dominant theme is

Do Not Cite, Quote or Distribute  18-63       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 accelerating the implementation of sustainability or low-carbon policies that target specific sectors or
 2 industries (Bhamidipati et al., 2019). For example, Altenburg and Rodrik (Altenburg and Rodrik, 2017)
 3 discuss green industrial polices including taxes, mandated technology phase outs, and the removal of
 4 subsidies as means of constraining polluting industries. Kivimaa et al. (Kivimaa and Martiskainen, 2018;
 5 Kivimaa et al., 2019a; Kivimaa et al., 2019b; Kivimaa et al., 2020) and Vihem鋕i et al. (2020) discuss low-
 6 carbon transitions in buildings, noting the important role that intermediaries play in facilitating policy
 7 reform. Nikulina et al. (2019) identify mechanisms for facilitating policy change in personal mobility
 8 including political leadership, combining carrots and sticks to incentivize behavioral change, and challenging
 9 current policy frameworks. These various examples reflect a fragmented approach to system transitions,
10 suggesting a large portfolio of such transition initiatives would be required to accelerate change or more
11 fundamental and cross-cutting policy drivers are needed (high agreement, limited evidence). Policies that
12 seek to promote social justice and equity, for example, could ultimately catalyze a broader range of
13 sustainability and climate actions than policies designed to address a specific sector or class of technology
14 (Delina and Sovacool, 2018; White, 2020).

15

16 In contrast with formal government policies, a second theme in accelerating transitions is that of civic
17 engagement (see also 18.4.3), which is reported to be an important opportunity for driving transitions
18 forward (high agreement, medium evidence). Ehnert et al. (2018) describe local organizations and civic
19 engagement in policy processes as an important engine for sustainability activities in European states.
20 Similarly, Ruggiero et al. (2021) note the potential to use civic organizations to appeal to local identities in
21 order to mobilize citizens to pursue energy transition initiatives among communities in the Baltic Sea region.
22 Gernert et al. (2018) attribute such influence to the ability of grassroots movements to bypass traditional
23 social and political norms and thereby experiment with new behaviors and processes. Moreover, civic
24 engagement is also the foundation for collective action including protest and civil disobedience (Welch and
25 Yates, 2018, Section 18.5.3.7). However, Haukkala (2018) observes that while green-transition coalitions in
26 Finland could be an agent of change driving energy transitions, the diversity of views among the various
27 grassroots actors could make consensus building difficult, thereby slowing transition initiatives.

28

29 A third theme is that of innovation, generally, and sustainability-oriented innovation, specifically (de Vries et
30 al., 2016; Geradts and Bocken, 2019; Loorbach et al., 2020), which creates opportunities for overcoming
31 existing transition barriers (very high confidence). For example, Valta (2020) describes the role of innovation
32 ecosystems � partnerships among companies, investors, governments, and academics � in accelerating
33 innovation (see also World Economic Forum, 2019). Burch et al. (Burch et al., 2016) describe the role of
34 small and medium-sized business entrepreneurship in promoting rapid innovation. Innovation extends
35 beyond pure technology considerations to consider innovation in practices and social organization (Li et al.,
36 2018; Psaltoglou and Calle, 2018; Repo and Matschoss, 2020). Zivkovic (2018), for example, discusses
37 "innovation labs" as accelerators for addressing so-called wicked problems like climate change through
38 multi-stakeholder groups. Meanwhile, Chaminade and Randelli (2020) describe a case study where structural
39 preconditions and place-based agency were important drivers of transitions to organic viticulture in Tuscany,
40 Italy.

41

42 The fourth theme is that of transition management (Goddard and Farrelly, 2018), particularly vis a vis,
43 disruptive technologies (I駃go and Albareda, 2016; Kuokkanen et al., 2019) or broader societal disruptions
44 (Brundiers, 2020; Davidsson, 2020; Hepburn et al., 2020; Schipper et al., 2020b). Recent literature has given
45 attention to how actors can use disruptive events, such as disasters, as a window-of-opportunity for
46 accelerating changes in policies, practices, and behaviors (high agreement, medium evidence) (Brundiers,
47 2018; Brundiers and Eakin, 2018). This is consistent with concepts in resilience thinking around `building
48 back better' after disasters (Fernandez and Ahmed, 2019). For example, Hepburn et al. discuss fiscal
49 recovery packages for COVID-19 as a means of accelerating climate action, with a particular influence on
50 clean physical infrastructure, building efficiency retrofits, investment in education and training, natural
51 capital investment, and clean research and development (Andrijevic et al., 2020b).

52

53

54 18.4 Agency and Empowerment for Climate Resilient Development

55

56 As reflected in the discussion of societal transitions (18.3), people and their values and choices play an
57 instrumental role in CRD. The agency of people to act on CRD is grounded in their worldviews, beliefs,

Do Not Cite, Quote or Distribute  18-64       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 values, and consciousness (Woiwode, 2020) and is shaped through social and political processes including
 2 how policies and decision-making recognize the voices, knowledges and rights of particular actors over
 3 others (very high confidence) (Harris and Clarke, 2017; Nightingale, 2017; Bond and Barth, 2020; Muok et
 4 al., 2021). Since the AR5, evidence on diverse forms of engagement by and among social, political and
 5 economic actors to support climate resilient development and sustainability outcomes, has increased. New
 6 forms of decision-making and engagement are emerging within the formal policy making and planning
 7 sphere, including co-production of knowledge, interventions grounded in the arts and humanities, civil
 8 participation and partnerships with business (Ziervogel et al., 2016a; Roberts et al., 2020). In addition, the
 9 set of actors that drive climate and development actions are recognized to extend beyond government and
10 formal policy actors to include civil society, education, industry, media, science and art (Ojwang et al., 2017;
11 Solecki et al., 2018; Heinrichs, 2020; Omukuti, 2020). This makes the power dynamics among actors and
12 institutions critical for understanding the role of actors in CRD (Buggy and McNamara, 2016; Camargo and
13 Ojeda, 2017; Silva Rodr韌uez de San Miguel, 2018).

14

15 The formal space for national, sub-national and international adaptation governance emerged at COP 16
16 (UNFCCC, 2010) when adaptation was recognized as a similar level of priority as greenhouse gas
17 mitigation. The Paris Agreement (UNFCCC, 2015) built on this and the 2030 Sustainable Development
18 Agenda (United Nations, 2015) to link adaptation to development and climate justice. It also highlighted the
19 importance of multi-level adaptation governance, including new non-state voices and climate actors that
20 widen the scope of adaptation governance beyond formal government institutions. For example, individuals
21 can act as agents of changes in their own behavior, such as via change in their consumption patterns, but also
22 generate change within organizations, fields of practice, and the political landscape of governance.
23 Accordingly, these interactions among actors across different scales implies the need for wider modes of,
24 and arena for, engagement around adaptation in order to accommodate a diversity of perspectives (high
25 agreement, medium evidence) (Chung Tiam Fook, 2017; Lesnikowski et al., 2017; IPCC, 2018a).

26

27 In most regions, such new institutional and informal arrangements are at an early stage of development (high
28 agreement, limited evidence). Further clarification and strengthening are needed to enable the fair sharing of
29 resources, responsibilities, and authorities to enable climate action to enable climate-resilient development
30 (Wood et al., 2017; IPCC, 2018a; Reckien et al., 2018). These are strongly linked to contested and
31 complementary worldviews of climate change and the actors that use these worldviews to justify, direct,
32 accelerate and deepen transformational adaptation and climate action.

33

34 18.4.1 Political Economy of Climate Resilient Development

35

36 Political economy studies (i.e., the origins, nature and distribution of wealth, and the ideologies, interests,
37 and institutions that shape it) explicitly addressing CRD are quite limited. Yet, there is an extensive post-
38 AR5 literature on political economy associated with various elements relevant to CRD including climate
39 change and development (Naess et al., 2015); vulnerability, adaptation, and climate risk (Sovacool et al.,
40 2015; Sovacool et al., 2017; Barnett, 2020); energy, decarbonization, and negative emissions technologies
41 (Kuzemko et al., 2019; Newell, 2019); degrowth and low-carbon economies (Perkins, 2019; Newell and
42 Lane, 2020); solar radiation management (Ott, 2018); planetary health and sustainability transitions and
43 transformation (Kohler et al., 2019) (Gill and Benatar, 2020).

44

45 Four key insights regarding the nexus of political economy and CRD emerge from this literature. First,
46 political economy drives coupled development-climate change trajectories and determines vulnerability,
47 thereby potentially subjecting those least responsible for climate change to the greatest risk (Sovacool et al.,
48 2015; Barnett, 2020). The prevailing political economy is itself now at risk as its legitimacy, viability and
49 sustainability are called into question (Barnett, 2020). Yet, as underpinning ideologies, interests and
50 institutions change, the drivers of vulnerability are often appropriated, the adaptation agenda is depoliticized,
51 and market-based solutions advocated (Barnett, 2020).

52

53 Second, assessment of this literature suggests four attributes of the political economy of adaptation influence
54 development trajectories in diverse settings, from Australia to Honduras and the Maldives (Sovacool et al.,
55 2015), as delivered through the Global Environment Facility's Least Developed Countries Fund (Sovacool et
56 al., 2017). These include enclosure (public resources or authority captured by private interests); exclusion
57 (stakeholders are marginalized from decision-making); encroachment (natural systems and ecosystem

Do Not Cite, Quote or Distribute  18-65       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 services compromised); and entrenchment (inequality exacerbated). These attributes hamper adaptation
 2 efforts, and reveal the political nature of adaptation (Dolsak and Prakash, 2018) and by extension CRD.
 3 Paradoxically, development initiatives labelled as `risk' reduction or resilience building or `equitable and
 4 environmentally sustainable', such as coastal restoration efforts in Louisiana, USA, can compound inequity
 5 and climate risk, and perpetuate unsustainable development (Gotham, 2016; Eriksen et al., 2021b).

 6

 7 Third, a long-held view is that the effects of mitigation are global while those of adaptation are local. A
 8 political economy perspective, however, underscores cross-scale linkages, and shows that local adaptation
 9 efforts, vulnerability and climate resilience are manifest in development trajectories that are shaped by both
10 local and trans-local drivers, and defined by unequal power relations that cross scales and levels (Sovacool et
11 al., 2015; Barnett, 2020; Newell, 2020), including in key sectors like energy (Baker et al., 2014) and
12 agriculture (Houser et al., 2019), as well as emergent blocs like BRICS (Power et al., 2016; Schmitz, 2017);
13 and sub-national constellations, like cities (Fragkias and Boone, 2016; B閚� et al., 2018).

14

15 Fourth, transitions towards CRD may be technically and economically feasible but are `saturated' with
16 power and politics (Tanner and Allouche, 2011) (18.3), necessitating focused attention to political barriers
17 and enablers of CRD (Newell, 2019). With a narrow window of time to contain dangerous levels of global
18 warming, political economy research calls for CRD trajectories that counter the globalized neoliberal
19 hegemony (Newell and Lane, 2020), especially given the pandemic, and the intersection of economic power
20 and public health, environmental quality, climate change, and human and indigenous rights (Bernauer and
21 Slowey, 2020; Schipper et al., 2020b).

22

23 Given these insights, CRD can be understood as the sum of complex multi-dimensional processes consisting
24 of large numbers of actions and social choices made by multiple actors from government, the private sector,
25 and civil society, with important influences by science and the media (very high confidence). These actions
26 and social choices are determined by the available solution space and options, along with a range of enabling
27 conditions (Section 18.4.2) that are largely bounded by individual and collective worldviews, and related
28 ethics and values. This view is consistent with sustainable development being a process constituted by
29 multiple actions that are contested and have path dependencies and context-sensitive synergies and trade-
30 offs with natural and embedded human systems as well as bounded by multiple and contested knowledges
31 and worldviews (Goldman et al., 2018; Heinrichs, 2020; Nightingale et al., 2020; Schipper et al., 2020b).

32

33 18.4.2 Enabling Conditions for Near-Term System Transitions

34

35 Given actors, institutions, and their engagement is fundamental to supporting system transitions needed for
36 CRD (18.3) this section assesses recent literature with respect to how the values, choices and behaviors of
37 those actors enable or constrain specific enabling conditions. Such enabling conditions represent
38 opportunities for policymakers to pursue actions that contribute to CRD beyond direct risk management
39 options such as climate adaptation and greenhouse gas mitigation (18.2.5.1, 18.2.5.2).

40

41 18.4.2.1 Governance and Policy

42

43 An overarching enabling conditions for achieving system transitions and transformations is the presence of
44 enabling governance systems (very high confidence). Recent literature on the translation of governance into
45 system transitions in practice suggests four key actions are important. The first is the critical reflection on so-
46 called `development solutions,' alternatively framed by some as `empty promises,' that worsen climate risk,
47 inequity, injustice and ultimately lead to unsustainable development (Mikulewicz, 2018; Mikulewicz and
48 Taylor, 2020). Examples include development aid (Scoville-Simonds et al., 2020), large-scale development
49 projects such as biofuel production in Ethiopia (Tufa et al., 2018), and urban growth management in
50 Vietnam (DiGregorio, 2015). The second is the recognition that while the power of different actors and
51 institutions is often tied to access to resources and the ability to constrain the actions of others, other
52 dimensions of power such as its ability to produce knowledge as well as its contingency on circumstances
53 and relationships are also important in enabling energy transitions: (Avelino et al., 2016; Avelino and
54 Wittmayer, 2016; Lockwood et al., 2016; Ahlborg, 2017; Avelino and Grin, 2017; Partzsch, 2017; Smith and
55 Stirling, 2018). Third, governance systems can help to develop productive interactions between formal
56 government institutions, the private sector, and civil society including the provision `safe arenas' for social
57 actors to deliberate and pursue transitional and transformational change (Haukkala, 2018; T鰎nberg, 2018;

Do Not Cite, Quote or Distribute  18-66       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 Strazds; Ferragina et al., 2020; Koch, 2020) (18.3.1, Box 18.1). Fourth, governance can address challenges
 2 such as climate change from a systems perspective and pursue interventions that address the interactions
 3 among development, climate change, equity and justice, and planetary health (Harvey et al., 2019; H鰈scher
 4 et al., 2019). This is evidenced by recent experience with the COVID-19 pandemic response as well as
 5 ongoing escalation of disaster risk associated with extreme weather events (Walch, 2019; Cohen, 2020;
 6 Schipper et al., 2020b; Wells et al., 2020).

 7

 8 One output from systems of governance is formal policy frameworks and policies that influence processes
 9 and outcomes of system transitions that support CRD (18.1.3). The Paris Agreement, for example, provides a
10 framework for CRD by defining a mitigation-centric goal of `limiting warming to well below 2癈 and
11 enabling a transition to 1.5癈' (UNFCCC, 2015). It also provides for a broadly defined global adaptation
12 goal (UNFCCC, 2015: Art. 7.1). The Nationally Determined Contributions (NDCs) are the core mechanism
13 for achieving and enhancing climate ambitions under the Paris Agreement. However, the pursuit of a given
14 NDC within a specific country will likely necessitate a range of other policy interventions that have more
15 immediate impact on technologies and behavior, implicating transitions in energy, industry, land, and
16 infrastructure (very high confidence (18.3.1). SDG-relevant activities are increasingly incorporated into
17 climate commitments in the NDCs (at last count 94 NDCs also addressed SDGs), contributing to several
18 (154 out of the 169) SDG targets (Brandi and Dzebo; Pauw et al., 2018). This reflects the potential of the
19 NDCs as near-term policy instruments and sign-posts for progress toward CRD (medium agreement, limited
20 evidence) (McCollum et al., 2018b).

21

22 As reflected by the SDGs (and SDG 13 specifically), the mainstreaming of climate change concerns into
23 development policies is one mechanism for pursuing sustainable development and CRD (very high
24 confidence). However, such mainstreaming has also been critiqued for perpetuating `development as usual',
25 reinforcing established development logics, structures and worldviews that are themselves contributing to
26 climate change and vulnerability (O'Brien et al., 2015) and for obscuring and depoliticizing adaptation
27 choices into technocratic choices (Murtinho, 2016; Webber and Donner, 2017; Benjaminsen and Kaarhus,
28 2018; Khatri, 2018; Scoville-Simonds et al., 2020). The coordinated implementation of sustainable
29 development policy and climate action is nonetheless crucial for ensuring that the attainment of one does not
30 come at the expense of others (Stafford-Smith et al., 2017). For example, aggressive pursuit of climate
31 policies that facilitate transitions in energy systems can undermine efforts to secure sustainability transitions
32 in other systems (18.3.1.1, 18.2.5.3, Table 18.7).

33

34 Several non-climate international policy agreements provide context for CRD such as the 1948 UN
35 Universal Declaration of Human Rights, the UN Declaration on the Rights of Indigenous Peoples (Hjerpe et
36 al., 2015); the Convention on Biological Diversity (CBD; UNFCCC, 1992) as well as the more recent Sendai
37 Framework for Disaster Risk Reduction (UNDRR, 2015) and the `new humanitarianisms' which seeks to
38 reduce the gap between emergency assistance and longer term development (Marin and Naess, 2017).
39 Collectively they provide a global policy framework that protects people's rights that are potentially
40 threatened by climate change (Olsson et al., 2014). These policies are relevant to transitions across multiple
41 systems, particular in societal systems toward more equitable and just development.

42

43 18.4.2.2 Economics and Sustainable Finance

44

45 18.4.2.2.1 Economics
46 System transitions toward CRD is contingent on reducing the costs of current climate variability on society
47 while making investments that prepare for the future effects of climate change. Climate change and
48 responses to climate change will affect many different economic sectors both directly and indirectly (Stern,
49 2007; IPCC, 2014a; Hilmi et al., 2017). As a consequence, the characteristics of economic systems will play
50 an important role in determining their resilience (very high confidence). These effects will occur within the
51 context of other developments, such as a growing world population, which increases environmental
52 pressures and pollution (Gonz醠ez-Hidalgo and Zografos, 2019; Gonz醠ez-Hidalgo and Zografos, 2020).
53 This impact is higher for developing countries than for high-income countries (Liobikien and Butkus,
54 2018). While looking for sustainable climate-resilient policies, many complex and interconnected systems,
55 including economic development, must be considered in the face of global-scale changes (Hilmi and Safa,
56 2010).

57

Do Not Cite, Quote or Distribute  18-67       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 Miller (2017) discusses some of the planning for, and application of, adaptation measures that improve
 2 sustainability noting the importance of considering a range of factors including complexities of
 3 interconnected systems, the inherent uncertainties associated with projections of climate change impacts, and
 4 the effects of global-scale changes such as technological and economic development for decision
 5 makers. For example, addressing climate impacts in isolation is unlikely to achieve equitable, efficient, or
 6 effective adaptation outcomes (very high confidence). Instead, integrating climate resilience into growth and
 7 development planning allows decision makers to identify what sustainable development policies can support
 8 climate resilient growth and poverty reduction and understand better how patterns and trends of economic
 9 development affect vulnerability and exposure to climate impacts across sectors and populations, including
10 distributional effects (Doczi, 2015). Markkanen and Anger-Kraavi (2019) highlighted that climate change
11 mitigation policy can influence inequality both positively and negatively. Although higher levels of poverty,
12 corruption and economic and social inequalities can increase the risk of negative outcomes, these potential
13 negative effects would be mitigated if inequality impacts were taken into consideration in all stages of policy
14 making (very high confidence).

15

16 The primary objective of economic and financial incentives around carbon emissions is to redirect
17 investment from high to low carbon technologies (Komendantova et al., 2016). Recent years have seen
18 policy interventions to incentivize transitions in energy, land, and industrial systems to address climate
19 change and sustainability focus on price-based, as opposed to quantity-based, interventions. Price-based
20 interventions aim at leveraging market mechanisms to achieve greater efficiency in the allocation of
21 resources and costs of mitigating climate change. For example, carbon pricing initiatives around the world
22 today cover approximately 8 gigatons of carbon dioxide emissions, equivalent to about 20% of global fossil
23 energy fuel emissions and 15% of total carbon dioxide greenhouse gas emissions (Boyce, 2018). Meanwhile,
24 environmental taxes and green public procurement push producers to eliminate the negative environmental
25 effects of production (Danilina and Trionfetti, 2019). There are several advantages for environmental
26 taxation including environmental effectiveness, economic efficiency, the ability to raise public revenue, and
27 transparency (very high confidence). These gains can provide more resource-efficient production
28 technologies and positively affect economic competitiveness (Costantini et al., 2018).

29

30 Policies encouraging eco-innovation, defined as "new ideas, behavior, products, and processes that
31 contribute to a decreased environmental burden" (Yurdakul and Kazan, 2020), can positively affect
32 economic competitiveness. By implementing policies to encourage eco-innovation, countries enhance their
33 energy efficiency. These gains can provide more resource-efficient production technologies and positively
34 affect economic competitiveness (very high confidence) (Liobikien and Butkus, 2018) (Costantini et al.,
35 2018). Other than eco-innovation, it is important to also consider exnovation, meaning the phasing out of old
36 technologies, as otherwise the expansion of supply could lead to a rebound due to cheaper prices for carbon-
37 based products (Arne Heyen et al., 2017; David, 2017). Hence, decarbonization strategies that set limits to
38 carbon-based trajectories can be beneficial. Quantity-based interventions--or so-called `command-and-
39 control' policies--involve constraints on the quantity of energy consumption or greenhouse gas emissions
40 through laws, regulations, standards and enforcement, with a focus on effectiveness rather than efficiency.

41

42 For a transition from dirty (more advanced) technologies to clean (less advanced) ones, market-based
43 instruments such as carbon taxes should be considered alongside subsidies and other incentives that
44 stimulate innovation (Acemoglu et al., 2016). Research and development in energy technologies, for
45 example, can help reduce costs of deployment and therefore the costs of operating in a carbon-constrained
46 world. H閙ous (2016) indicates that a unilateral environmental policy which includes both clean research
47 subsidies and trade tax can ensure sustainable growth, but unilateral carbon taxes alone might increase
48 innovation in polluting sectors and would not generally lead to sustainable growth.

49

50 18.4.2.2.2 Climate finance
51 Achieving progress on system transitions will be contingent on the ability of actors and institutions to access
52 the financing they need to invest in innovation, adaptation and mitigation, and broader system change (very
53 high confidence). By greening their investment portfolios, investors can support reduction in vulnerability to
54 the consequences of climate change and the reduction of greenhouse gas emissions. Finance can contribute
55 to the reduction of GHG emissions, for example, by efficiently pricing the social cost of carbon, by
56 reflecting the transition risks in the valuation of financial assets, and by channeling investments in low-
57 carbon technologies (OECD, 2017). At the same time, there is a growing need to spur greater public and

Do Not Cite, Quote or Distribute  18-68       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 private capital into climate adaptation and resilience including climate-resilient infrastructure and nature-
 2 based solutions to climate change. For instance, the Green Climate Fund, established within the framework
 3 of the UNFCCC, is assisting developing countries in adaptation and mitigation initiatives to counter climate
 4 change.

 5

 6 Recent evidence sheds light on the magnitude and pervasiveness of climate risk exposure for global banks
 7 and financial institutions. According to Dietz et al. (2016), up to about 17% of global financial assets are
 8 directly exposed to climate risks, particularly the impacts of extreme weather events on assets and their
 9 outputs. However, when indirect exposures via financial counterparts are considered, the share of assets
10 subject to climate risks is much larger (40-54%) (Battiston et al., 2017). Hence, the magnitude of climate-
11 change-related risks is substantial, and similar to the ones that started the 2008 financial crisis (high
12 agreement, limited evidence).

13

14 Financial actors increasingly recognize that the generation of long-term, sustainable financial returns is
15 dependent on a stable, well-functioning and well-governed social, environmental and economic systems
16 (very high confidence) (Shiller, 2012; Schoenmaker and Schramade, 2020). Institutional approaches to a
17 variety of environmental domains (Krueger et al., 2019), which seek to integrate the pursuit of green
18 strategies with financial returns include targeted investments in green assets (e.g., green bonds, clean energy
19 public equity) and specialized funds/vehicles for as renewable energy infrastructure (Tolliver et al., 2019;
20 Gibon et al., 2020); cleantech venture capital and alternative finance (Gianfrate and Peri, 2019); investment
21 screening to steer capital to green industries (Nielsen and Skov, 2019; Ambrosio et al., 2020); and active
22 ownership to influence organizational behavior (Silvola and Landau, 2021).

23

24 Despite the expansion of green mandates across the investment chain, definitions of some of the asset classes
25 associated with green investing are ambiguous and poorly defined. The EU taxonomy for sustainable
26 activities is a promising step in the right direction. For example, a "green" label for bonds is often stretched
27 to encompass financing facilities of issuers that misrepresent the actual environmental footprint of their
28 operations (the so-called risk of "greenwashing"). Even in cases where the bonds' proceeds are actually used
29 to finance green projects, investors often remain exposed to both the green and "brown" assets of the issuers
30 (Gianfrate and Peri, 2019; Flammer, 2020). The heterogeneity of metrics and rating methodologies (along
31 with inherent conflict of interests between issuers, investors and score/rating providers) results in
32 inconsistent and unreliable quantification of the actual environmental footprint of corporate and sovereign
33 issuers (Battiston et al., 2017; Busch et al.).

34

35 In order to promote financial climate-related disclosures for companies and financial intermediaries, the
36 financial system could play a key role in pricing carbon and in allocating capital toward low-carbon emission
37 companies (Aldy and Gianfrate, 2019; Bento and Gianfrate, 2020; Aldy et al., 2021). Stable and predictable
38 carbon-pricing regimes would significantly contribute to fostering financial innovation that can help further
39 accelerate the decarbonization of the global economy even in jurisdictions which are more lenient in
40 implementing climate mitigation actions (very high confidence) (Baranzini et al., 2017). A growing number
41 of financial regulators are intensifying efforts to enhance climate-related disclosure of financial actors. In
42 particular, the Financial Stability Board created the Task Force on Climate-related Financial Disclosures
43 (TCFD) to improve and increase reporting of climate-related financial information. Several countries are
44 considering implementing mandatory climate risk disclosure in line with TCFD's recommendations. Central
45 Banks are also considering mandatory disclosure and climate stress-testing for banks. For instance, in
46 November 2020 the European Central Bank (ECB) published a guide on climate-related and environmental
47 risks explaining how the ECB expects banks to prudently manage and transparently disclose such risks under
48 current prudential rules. The ECB also announced that banks in the Euro-zone will be stress tested on their
49 ability to withstand climate change related risks. In addition to disclosure requirements and stress-testing,
50 some Central Banks are considering the possibility of steering or tilting the allocation of their assets to favor
51 the less polluting issuers (Schoenmaker, 2019). This, in turn, would translate into lower cost of capital for
52 cleaner sectors, significantly accelerating the greening of the real economy.

53

54

55 [START BOX 18.7 HERE]

56

Do Not Cite, Quote or Distribute  18-69       Total pages: 197
    FINAL DRAFT                        Chapter 18          IPCC WGII Sixth Assessment Report

1 Box 18.7: `Green' Strategies of Institutional Investors

 2

 3 Negative and positive screening. Investors assess the carbon footprint of issuers and identify the best and
 4 worst performers (Boermans and Galema, 2019). The issuers with excessive carbon footprint are divested
 5 and fall into the "exclusion lists" (negative screening). Alternatively, the investors commit to pick only the
 6 best in class (positive screening). As a bare minimum, screening approaches force more transparent
 7 environmental reporting from issuers. In the most optimistic scenario, in order to avoid exclusion lists issuers
 8 may progressively divest their non-green operations. In the long term, the combination of positive and
 9 negative screening will reward sustainable issuers relative to non-green sectors, thus reducing the cost of
10 capital for less polluting entities.

11

12 Active ownership. Equity investors can exercise the voting rights at shareholders' meetings in relation to
13 governance and business strategy, including the environmental performance. In addition, institutional
14 investors engage with the management and the boards of directors of investee companies. Active ownership
15 is therefore defined as the full exercise of the rights that accrue to the "owners" of the securities issued by
16 companies (Dimson et al., 2015; Dimson et al., 2020). Active owners are entitled to question and challenge
17 the robustness of financial analyses and the risk assessment behind strategic decisions including the
18 environmental footprint ones. For instance, since fossil fuel businesses face the prospect of dramatic
19 business decline (Ansar et al., 2013) and must revisit their business model to survive, active ownership by
20 institutional investors may foster the transition to cleaner production and supply chain. Companies more
21 exposed to carbon risks particularly need the active support of long-term shareholders. In turn, investors
22 adopting an active ownership approach can manage their holdings' exposure to climate change risks, thus
23 protecting the value of their investments on a long-term horizon (Krueger et al., 2019).

24

25 Specialized financial instruments and investors. New asset classes have been created to address the climate
26 change challenge. Also specialized investment funds and vehicles came to life with the primary objective of
27 addressing climate issues. While these financial instruments and funds prioritize the achievement of climate
28 objectives, they do not sacrifice financial returns and are able to attract private capital. To mention a few
29 examples:

30

31  Green bonds are typically issued by companies, banks, municipalities, and governments with the

32  commitment to use the proceeds exclusively to finance or refinance green projects, assets or business

33  activities. These bonds are equivalent to any other bond issued by the same entity except for the label of

34  "greenness" that ideally is verified ex-ante at the launch and ex-post when the proceeds are actually used

35  by the issuer. Early evidence show that green bonds do not penalize financially issuers (Gianfrate and

36  Peri, 2019; Flammer, 2020).

37  Carbon funds are designed to help countries achieve long-term sustainability typically financing forest

38  conservation. They are intended to reduce climate change impacts from forest loss and degradation.

39  Project finance. New renewable energy initiatives are likely to recur more and more to project finance.

40  Project finance relies on the creation of a special purpose vehicle (SPV), which is legally and

41  commercially self-contained and serves only to run the renewable energy project. The SPV is financed

42  without (or very limited) guarantees from the sponsors (typically energy companies: investors are

43  therefore paid back on the basis only of SPV's future cash flows only and cannot recourse on the

44  sponsors' assets (Steffen, 2018).

45  Cleantech venture capital. These funds invest exclusively in early-stage companies working on

46  innovative but not yet fully tested clean technologies. The risk profile of such investments is usually

47  very high. The extent to which this segment of the financial industry can successfully support "deep"

48  energy innovations is still debated (Gaddy et al., 2017). When cleantech start-ups develop hardware

49  requiring a high upfront investment, support from the public sector seems necessary in order to attract

50  further investments from large corporations and patient institutional investors.

51  Crowdfunding and alternative finance are emerging as a channel to both finance small-scale clean

52  energy projects as well as fund early stage innovative clean technologies (Cumming et al., 2017; Bento

53  et al., 2019).

54

55 [END BOX 18.7 HERE]

56

57

    Do Not Cite, Quote or Distribute   18-70                                          Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 18.4.2.3 Institutional capacity

 2

 3 Institutional capacity for system transitions refers to the capacity of structures and processes, rules, norms,
 4 and cultures to shape development expectations and actions aimed at durable improvements in human well-
 5 being. The AR5 highlighted the need for strong institutions to create enabling environments for adaptation
 6 and greenhouse gas mitigation action (Denton et al., 2014). Institutions stand within the social and political
 7 practices and broader systems of governance that ultimately drive adaptation and development processes and
 8 outcomes. They are thus produced by them and can become tools by which some actors constrain the actions
 9 of others (Gebreyes, 2018). As a consequence, they and can become a significant barrier to change, whether
10 incremental or more transformational (very high confidence). The post-AR5 focus on transformational
11 adaptation and resilience present in the literature suggests that institutions that enable system transitions
12 toward CRD are secure enough to facilitate a wide range of voices, and legitimate enough to change goals or
13 processes over time, without reducing confidence in their efficacy.

14

15 The limited literature on institutions and pathways relevant to system transitions and CRD suggests that
16 institutions are most effective when taking a development-first approach to adaptation. This is consistent
17 with the principles of CRD which emphasizes not simply reducing climate risk, but rather making
18 development processes resilient to the changing climate. There is agreement in this literature that such an
19 approach allows for the effective integration of climate challenges into existing policy and planning
20 processes (very high confidence) (Pervin et al., 2013; Kim et al., 2017b; Mogelgaard et al., 2018). However,
21 this approach generally rests on an incremental framing of institutional change (Mahoney and Thelen, 2009)
22 based on two critical assumptions. The first is that existing processes and institutions are capable of bringing
23 about system transitions that generate desired development outcomes and thus can be considered appropriate
24 vehicles for the achievement of CRD. A large critical literature questions the efficacy of formal state and
25 multilateral institutions. The evidence for the ability of local, informal institutions to achieve development
26 goals remains uneven, with robust evidence of positive impacts on public service delivery, but more
27 ambiguous evidence on behavior changes associated with strengthened institutions (Berkhout et al., 2018).
28 The second is that the mainstreaming of adaptation will bring about changes to currently unsustainable
29 development practices and pathways, instead of merely strengthening development-as-usual by subsuming
30 adaptation to existing development pathways and allowing them to endure in the face of growing stresses
31 (Eriksen et al., 2015; Godfrey-Wood and Otto Naess, 2016; Scoville-Simonds et al., 2020). There is
32 evidence that countries with poor governance have limited adaptation planning or action at the national level,
33 even when other determinants of adaptive capacity are present (Berrang-Ford et al., 2014). This suggests
34 that, in these contexts, adaptation efforts are likely to be subsumed to existing government goals and actions,
35 rather than having transformational impact.

36

37 18.4.2.4 Science, Technology & Innovation

38

39 Ongoing innovations in technology, finance, and policy have enabled more ambitious climate action over the
40 past decade, including significant growth in renewable energy, electrical vehicles, and energy efficiency.
41 However, access to, and the benefits of, that innovation have not been evenly distributed among global
42 regions and communities and continued innovation is needed to facilitate climate action and sustainable
43 development (very high confidence). Policymakers need useful science and information (Kirchhoff et al.,
44 2013; Calkins, 2015; IPCC, 2019f) to make informed decisions about possible risks, and the benefits, costs,
45 and trade-offs of available adaptation, mitigation, and sustainable development solutions (i.e., Article 4.1 of
46 the Paris Agreement; UNFCCC, 2015). Moreover, recent literature has emphasized the need for deep
47 technological, as well social, changes to avert the risks of conventional development trajectories (Gerst et al.,
48 2013; IPCC, 2014a).

49

50 An effective and innovative technological regime is one that is integrated with local social entities across
51 different modes of life, local governance processes (Pereira, 2018; Nightingale et al., 2020); and local
52 knowledge(s), which increasingly support adaptation to socio-environmental drivers of vulnerability
53 (Schipper et al., 2014; Nalau et al., 2018; IPCC, 2019f). These actors and their knowledge are often ignored
54 in favor of knowledge held by experts and policymakers, exacerbating uneven power relations (Naess, 2013;
55 Nightingale et al., 2020). For example, achieving sustainability and shifting towards a low carbon energy
56 system (e.g., hydropower dams, wind farms) remains a contested space with divergent interests, values and
57 prospects of future (Bradley and Hedr閚, 2014; Avila, 2018; Mikulewicz, 2019), and potential impacts on

Do Not Cite, Quote or Distribute  18-71       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 human rights as embodied by the Paris Agreement (UNFCCC, 2015). A number of studies have emphasized
 2 the limits of relying upon technology innovation and deployment (e.g., expansion of renewable energy
 3 systems and/or carbon capture) as a solution to challenges of climate change and sustainable development
 4 (18.3.1.2). This is because such solutions may fail to consider the local historical contexts and barriers to
 5 participation of vulnerable communities, restricting their access to land, food, energy, and resources for their
 6 livelihoods.

 7

 8 18.4.2.5 Monitoring and Evaluation Frameworks

 9

10 Enabling system transitions toward CRD is dependent in part on the ability to monitor and evaluate system
11 transitions and broader development pathways to identify effective interventions and barriers to their
12 implementation (very high confidence). However, the monitoring and evaluation of individual system
13 transitions, much less CRD, remains highly challenging for multiple reasons (Persson, 2019). The highly
14 contextual nature of resilience, adaptation and sustainable development means that, unlike climate
15 mitigation, it is difficult to define universal metrics or targets for adaptation and resilience (Pringle and
16 Leiter, 2018), (Brooks et al., 2014). This is demonstrated by the Paris Agreement's global goal for
17 adaptation, The mismatch between timescales associated with resilience and adaptation interventions and
18 those over which the results of such interventions are expected to become apparent tends to result in a focus
19 on the measurement of spending, outputs, and short-term outcomes, rather than longer-term impacts (Brooks
20 et al., 2014; Pringle and Leiter, 2018). The need to assess resilience and adaptation against a background of
21 evolving climate hazards, and to link resilience and adaptation with development outcomes, present further
22 methodological challenges (very high confidence) (Brooks et al., 2014).

23

24 Currently, the ability to monitor different components of CRD are in various stages of maturity (very high
25 confidence). Monitoring of the sustainable development goals, for example, is a routine established practice
26 at global and regional levels, and UNDP publishes annual updates on progress toward the SDGs (United
27 Nations, 2021). For resilience, Brooks et al. (2014) identify three broad approaches to its measurement, each
28 of which could offer potential mechanisms for monitoring progress toward CRD. One is a `hazards'
29 approach, in which resilience is described in terms of the magnitude of a particular hazard that can be
30 accommodated by a system, useful in contexts where thresholds in climate and related parameters can be
31 identified and linked with adverse impacts on human populations, infrastructure and other systems (Naylor et
32 al., 2020). An `impacts' approach is one in which resilience is measured in terms of actual or avoided
33 impacts and is suited for tracking adaptation success in delivering CRD over longer timescales, for example
34 at the national level (Brooks et al., 2014). Finally, a `systems' approach is one where resilience is described
35 in terms of the characteristics of a system using quantitative or qualitative indicators which are often
36 associated with different `dimensions' of resilience (Serfilippi and Ramnath, 2018; Saja et al., 2019). This
37 allows measurement of key indicators that are proxies for resilience at regular intervals, even in the absence
38 of significant climate hazards and associated disruptions (very high confidence) (Brooks et al., 2014) (see
39 also Cross-Chapter Box ADAPT in Chapter 1). Similar criteria could be applied to evaluating adaptation
40 options and their implementation as well as various interventions in pursuit of SDGs.

41

42 18.4.3 Arenas of Engagement

43

44 Much of the enabling conditions for system transitions discussed in 18.4.2 are inherently linked to actors and
45 their agency in pursuing system change. Yet, a significant literature has developed since the AR5 exploring
46 note only the role of different actors in pursuing adaptation, mitigation, and sustainable development options,
47 but also how those actors interact with one another to drive outcomes. CRD pathways are determined by the
48 interactions between societal actors and networks, including government, civil society and the private sector,
49 as well as science and the media. The resultant social choices and cumulative private and public actions (and
50 inactions) are institutionalized through both formal and informal institutions that evolve over time and seek
51 to provide societal stability in the face of change. The degree to which the emergent pathways foster just and
52 climate resilient development depends on how contending societal interests, values and worldviews are
53 reconciled through these interactions. These interactions occur in many different arenas of engagement, i.e.,
54 the settings, places and spaces in which societal actors interact to influence the nature and course of
55 development, including political, economic, socio-cultural, ecological, knowledge-technology and
56 community arenas (Figures 18.1, 18.2).

57

Do Not Cite, Quote or Distribute  18-72       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 For example, political arenas range from formalized election and voting procedures to more informal and
 2 less transparent practices, like special interest lobbying. Town squares and streets can become sites of
 3 political struggle and dissent, including protests against climate inaction. As a more specific case-in-point,
 4 the formal space for national, sub-national and international adaptation governance emerged at COP 16
 5 (UNFCCC, 2010) when adaptation was recognized as having a similar level of priority as mitigation. The
 6 Paris Agreement (UNFCCC, 2015) built on this and the 2030 Sustainable Development Agenda (United
 7 Nations, 2015) to link adaptation to development and climate justice, widening the scope of adaptation
 8 governance beyond formal government institutions. It also highlighted the importance of multi-level
 9 adaptation governance, including non-state voices from civil society and the private sector. This implied the
10 need for wider arenas and modes of engagement around adaptation (Chung Tiam Fook, 2017; Lesnikowski
11 et al., 2017; IPCC, 2018a) that facilitate coordination and convergence among these diverse actors including
12 individual citizens to collectively solve problems and unlock the synergies between adaptation and
13 mitigation and sustainable development (IPCC, 2018a; Romero-Lankao et al., 2018).

14

15 There are many other visible and less visible arenas of engagement in the other interconnected spheres of
16 societal interaction spanning scales from the local to international level. The metaphor of arenas derives
17 from diverse social and political theory, with applications in studies of, among other things, governance
18 transformation and transitions (Healey, 2006; J鴕gensen, 2012; J鴕gensen et al., 2017). It underscores that
19 these arenas can be enduring or temporary in nature, are historically situated and often spatially bounded,
20 and signifies the many different mechanisms by which societal actors interact in dynamic and emergent
21 ways. Power and politics impact access and influence in these arenas of engagement � with varying levels of
22 inclusion and exclusion shaping the nature and trajectory of development. In practice, some arenas of
23 engagement are `struggle arenas' as different societal actors strive to influence the trajectory of development,
24 with inevitable winners and losers.

25

26 Institutional arrangements to foster CRD are at an early stage of development in most regions (medium
27 agreement, limited evidence). They need to be further clarified and strengthened to enable a sharing of
28 resources and responsibilities that facilitate climate actions embracing climate resilience, equity, justice,
29 poverty alleviation and sustainable development (Wood et al., 2017; IPCC, 2018a; Reckien et al., 2018).
30 These endeavours are strongly influenced by how contested and complementary worldviews about climate
31 change and development are mobilised by societal actors to justify, direct, accelerate and deepen
32 transformational climate action or entrench maladaptive business as usual practices (18.4.3.1).

33

34 18.4.3.1 Worldviews

35

36 Worldviews are overarching systems of meaning and meaning-making that inform how people interpret,
37 enact, and co-create reality (De Witt et al., 2016). Worldviews shape the vision, beliefs, attitudes, values,
38 emotions, actions, and even political and institutional arrangements. As such, they can promote holistic,
39 egalitarian approaches to enable, accelerate and deepen climate action and environmental care (Ramkissoon
40 and Smith, 2014; De Witt et al., 2016; Lacroix and Gifford, 2017; Sanganyado et al., 2018; Brink and
41 Wamsler, 2019). Alternatively, they can also serve as significant barriers to system transitions and
42 transformation, based on anthropocentric, mechanistic and materialistic, worldviews and the utilitarian,
43 individualist or skeptical values and attitudes they often promote (very high confidence) (Beddoe et al., 2009;
44 van Egmond and de Vries, 2011; Stevenson et al., 2014; Zummo et al., 2020).

45

46 Traditional, modern and postmodern worldviews have different, and in many ways, complementary
47 potentials for integrative diverse approaches to climate action and sustainable development. They can also
48 destabilize climate-sensitive societal values (van Egmond and de Vries, 2011; Van Opstal and Hug�, 2013;
49 De Witt et al., 2016; Shaw, 2016) which are predictors of concern (Shi et al., 2015). Among the challenges
50 of strongly different climate-related worldviews, is that they rarely co-exist. Some worldviews become
51 incompatible or hostile to other worldviews, openly seeking to dominate, eliminate or segregate competing
52 perspectives (medium agreement, medium evidence) (de Witt, 2015; Jackson, 2016; Nightingale, 2016; Xue
53 et al., 2016; Goldman et al., 2018).

54

55 To address these difficult contests, climate- and global environmental change-related worldviews are often
56 scientized. This can exclude other worldviews which ultimately narrows understanding of climate change
57 and the solution space. Hence, the post-AR5 literature on worldviews focuses on the numerous meanings,

Do Not Cite, Quote or Distribute  18-73       Total pages: 197
    FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1 associations, narratives and frames of climate change and how these shape perceptions, attitudes and values
2 (Morton, 2013; Boulton, 2016; Hulme, 2018; Nightingale B鰄ler, 2019). The recognition of the diversity of
3 interpretations and meanings has led to multidisciplinary and transdisciplinary research that incorporates the
4 humanities and the arts (Murphy, 2011; Elliott and Cullis, 2017; Steelman et al., 2019; Tauginien et al.,
5 2020), feminist studies (MacGregor, 2003; Demeritt et al., 2011; Bell, 2013; Brink and Wamsler, 2019;
6 Plesa, 2019) and religious studies (Sachdeva, 2016; McPhetres and Zuckerman, 2018) to examine diverse
7 understandings of reality and knowledge possibilities around climate change. In addition, literature on
8 cultural cognition, epistemological plurality and relational ontologies draws on non-Western worldviews and
9 forms of knowledge (Goldman et al., 2018) (Jackson, 2016; Nightingale, 2016; Xue et al., 2016).

10

11 On the other hand, the tendence for certain worldviews to dominate the policy discourse has the potential to
12 exacerbate social, economic and political inequities (very high confidence). ontological, epistemic and
13 procedural injustices. Research aimed at exploring the existing political ontology and knowledge politics of
14 exclusion that marginalize certain communities and actors originated in academic, or scientific perspectives.

15 This includes institutions such as the IPCC and is subsequently replicated in social representations, including
16 the media, public policy and the development agenda, narrowing possibilities for social transformation
17 (Jackson, 2014; Luton, 2015; Escobar, 2016; Burman, 2017; Newman et al., 2018; Sanganyado et al., 2018;
18 Wilson and Inkster, 2018).

19

20

21 [START CROSS-CHAPTER BOX INDIG HERE]

22

23 Cross-Chapter Box INDIG: The Role of Indigenous Knowledge and Local Knowledge in

24  Understanding and Adapting to Climate Change

25

26 Authors: Tero Mustonen (Finland), Sherilee Harper (Canada), Gretta Pecl (Australia), Vanesa Cast醤 Broto

27 (Spain), Nina Lansbury (Australia), Andrew Okem (Nigeria/South Africa), Ayansina Ayanlade (Nigeria),

28 Jackie Dawson (Canada), Pauline Harris (Aotearoa-New Zealand), Pauliina Feodoroff (Finland), Deborah
29 McGregor (Canada)

30

31 Indigenous knowledge refers to the understandings, skills and philosophies developed by societies with long

32 histories of interaction with their natural surroundings (UNESCO, 2018; IPCC, 2019a). Local knowledge
33 refers to the understandings and skills developed by individuals and populations, specific to the places where
34 they live (UNESCO, 2018; IPCC, 2019a). Indigenous knowledge and local knowledge are inherently
35 valuable but have only recently begun to be appreciated and in western scientific assessment processes in
36 their own right (Ford et al., 2016). In the past these often endangered ways of knowing have been suppressed
37 or attacked (Mustonen, 2014). Yet these knowledge systems represent a range of cultural practices, wisdom,
38 traditions, and ways of knowing the world that provide accurate and useful climate change information,
39 observations, and solutions (very high confidence) (Table Cross-Chapter Box INDIG.1). Rooted in their own
40 contextual and relative embedded locations, some of these knowledges represent unbroken engagement with
41 the earth, nature and weather for many tens of thousands of years, with an understanding of the ecosystem
42 and climatic changes over longer-term timescales that is held both as knowledge by Indigenous Peoples and
43 Local Peoples as well as in the archaeological record (Barnhardt and Angayuqaq, 2005; UNESCO, 2018).

44

45 Indigenous Peoples around the world often hold unique worldviews that link today's generations with past
46 generations. In particular, many Indigenous Peoples consider concepts of responsibility through
47 intergenerational equity, thereby honouring both past and future generations (Matsui, 2015; McGregor et al.,

48 2020). This can often be in sharp contrast to environmental valuing and decision-making that occurs in

49 Western societies (Barnhardt and Angayuqaq, 2005). Therefore, consideration of Indigenous knowledge and
50 local knowledge needs to be a priority in the assessment of adaptation futures (Nakashima et al., 2012)(Ford
51 et al., 2016) (Chapter 1), although adequate Indigenous cultural and intellectual property rights require legal
52 and non-legal measures for recognition and protection (Janke, 2018).

53

54 Indigenous knowledge and local knowledge are crucial to address environmental impacts, such as climate
55 change, where the uncertainty of outcome is high and a range of responses are required (Mackey and
56 Claudie, 2015). However, working with this knowledge in an appropriate and ethically acceptable way can
57 be challenging. For instance, questions of data `validity' and the requirement to communicate such

    Do Not Cite, Quote or Distribute     18-74                                       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 knowledge in the dominant language can lead to inaccurate portrayals of Indigenous knowledge as inferior to
 2 science. This may overlook the uniqueness of Indigenous knowledge and then lead to the overall devaluation
 3 of Indigenous political economies, cultural ecologies, languages, educational systems, and spiritual practices
 4 (Smith, 2013; Sillitoe, 2016; Naude, 2019; Barker and Pickerill, 2020). Furthermore, Indigenous knowledge
 5 is too often only sought superficially � focusing only on the `what', rather than the `how' of climate change
 6 adaptation and/or seen through the lenses of `romantic glorification' leaving little room for the knowledge to
 7 be expressed as authored by the communities and knowledge holders themselves (Yunkaporta, 2019).

 8

 9 Multiple knowledge systems and frameworks

10

11 Indigenous knowledge systems include not only the specific narratives and practices to make sense of the
12 world, but also profound sources of ethics and wisdom. They are networks of actors and institutions that
13 organise the production, transfer and use of knowledge (L鰂marck and Lidskog, 2017). There is a pluralism
14 of forms of knowledge that emerge from oral traditions, local engagement with multiple spaces, and
15 Indigenous cultures (Peterson et al., 2018). Recognising such multiplicity of forms of knowledge has long
16 been an important concern within sustainability science (Folke et al., 2016). Less dominant forms of
17 knowledge should not be put aside because they are not comparable or complementary with scientific
18 knowledge (Brattland and Mustonen, 2018; Mustonen, 2018; Ford et al., 2020; Ogar et al., 2020). Instead,
19 Indigenous knowledge and local knowledge can shape how climate change risk is understood and
20 experienced, the possibility of developing climate change solutions grounded in place-based experiences,
21 and the development of governance systems that match the expectations of different Indigenous knowledge
22 and local knowledge holders (very high confidence).

23

24 Different frameworks that enable the inclusion of Indigenous knowledge have emerged from efforts to utilise
25 more than one knowledge system (high evidence, high agreement). For example, the Intergovernmental
26 Science-Policy Platform on Biodiversity and Ecosystem Services (IPBES) has developed a `nature's
27 contribution to peoples' framework that provides a common conceptual vocabulary and structural analysis
28 (D韆z et al., 2015; Teng� et al., 2017; D韆z et al., 2018; Peterson et al., 2018). The IPBES approach
29 complements other efforts to study areas of intersection between scientific and Indigenous worldviews
30 (Barnhardt and Angayuqaq, 2005; Huaman and Sriraman, 2015) or `boundaries' that illustrate `blind spots'
31 in scientific knowledge (Cash et al., 2003; Clark et al., 2016; Brattland and Mustonen, 2018). These
32 frameworks highlight areas of collaboration but provide less guidance in areas where sources of evidence
33 conflict across different knowledge systems (L鰂marck and Lidskog, 2017). These experiences suggest that
34 the inclusion of Indigenous knowledge and local knowledge in international assessments may transform the
35 process of assessment of scientific, technical, and socio-economic evidence (medium evidence, high
36 agreement). These knowledge systems also point to novel discoveries that may be still unknown to the
37 scientific world but have been known by communities for millennia (Mustonen and Feodoroff, 2020).

38

39 The importance of free and prior-informed consent

40

41 Obtaining free and prior-informed consent is a necessary but not sufficient condition to engage in knowledge
42 production with Indigenous Peoples (Sillitoe, 2016). Self-determination in climate change assessment,
43 response, and governance is critical (Chakraborty and Sherpa, 2021), and Indigenous Peoples are actively
44 contributing to respond to climate change (Etchart, 2017). Climate change assessment and adaptation should
45 be self-determined and led by Indigenous Peoples, acknowledge the importance of developing genuine
46 partnerships, respect Indigenous knowledge and ways of knowing, and acknowledge Indigenous Peoples as
47 stewards of their environment (Country et al., 2016; Country et al., 2018; ITK, 2019; Barker and Pickerill,
48 2020; Chakraborty and Sherpa, 2021). Supporting Indigenous Peoples' leadership and rights in climate
49 adaptation options at the local, regional, national and international levels is an effective way to ensure that
50 such options are adapted to their living conditions and do not pose additional detrimental impacts to their
51 lives (very high confidence). Chapter 18 shows that the transformations required to deliver climate resilient
52 futures will create societal disruptions, with impacts that are most often unevenly experienced by groups
53 with high exposure and sensitivity to climate change, including Indigenous Peoples and local communities
54 (Schipper et al., 2020a). Climate-resilient futures depend on finding strategies to address the causes and
55 drivers of deep inequities (Chapter 18). For example, climate resilient futures will depend on recognising the
56 socio-economic, political and health inequities that often affect Indigenous Peoples (Mapfumo et al., 2016;
57 Ludwig and Poliseli, 2018) (very high confidence).

Do Not Cite, Quote or Distribute  18-75       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1

 2 International conventions to support and utilize Indigenous knowledge and local knowledge

 3

 4 Several tools within international conventions may support instruments to develop equitable processes that
 5 facilitate the inclusion Indigenous knowledge and leadership in climate change adaptation initiatives. The
 6 International Labour Convention 69 recognised Indigenous People's right to self-determination in 1989
 7 (ILO, 1989). The United Nations' Declaration on the Rights of Indigenous Peoples (United Nations, 2007)
 8 includes articles on the right to development (Article 23), the right to maintain and strengthen their
 9 distinctive spiritual relationship and to uphold responsibilities to future generations (Article 25), and the right
10 to the conservation and protection of the environment and the productive capacity of their territories (Article
11 29). Article 26 upholds the right to the lands, territories and resources, the right to own, use, develop and
12 control the lands, and legal recognition and protection of these lands, territories, and resources. Indigenous
13 Peoples are also recognized within the Sustainable Development Goals as a priority group (Carino and
14 Tamayo, 2019). International events such as the `Resilience in a time of uncertainty: Indigenous Peoples and
15 Climate Change' Conference brought together Indigenous Peoples' representatives and government leaders
16 from around the world to discuss the role of Indigenous Peoples in climate adaptation (UNESCO, 2015).

17

18 The value of Indigenous knowledge and local knowledge in climate adaptation planning

19

20 There have been increasing efforts to enable Indigenous knowledge holders to participate directly in IPCC
21 assessment reports (Ford et al., 2012; Nakashima et al., 2012; Ford et al., 2016). Adaptation efforts have
22 benefited from the inclusion of Indigenous knowledge and local knowledge (IPCC, 2019e) (very high
23 confidence). Moreover, it has been recognized that including Indigenous knowledge and local knowledge in
24 IPCC reports can contribute to overcoming the combined challenges of climate change, food security,
25 biodiversity conservation, and combating desertification and land degradation (IPCC, 2019c) (high
26 confidence). Limiting warming to 1.5癈 necessitates building the capability of formal assessment processes
27 to respect, include and utilize Indigenous knowledge and local knowledge (IPCC, 2018a) (medium evidence,
28 high agreement).

29

30 However, these efforts have been accompanied by a recognition that `integration' of Indigenous knowledge
31 and local knowledge cannot mean that those knowledge systems are subsumed or required to be validated
32 through typical scientific means (Gratani et al., 2011; Matsui, 2015). Such a critique of `validity' can be
33 inappropriate, unnecessary, can disrespect Indigenous Peoples' own identities and histories, limits the
34 advancement and sharing of these perspectives in the formal literature, and overlooks the structural drivers
35 of oppression and endangerment that are associated with Western civilization (Ford et al., 2016). Moreover,
36 by underutilizing Indigenous knowledge and local knowledge systems, opportunities that could otherwise
37 facilitate effective and feasible adaptation action can be overlooked. We should also reserve space for the
38 understanding that each cultural knowledge system, building on linguistic-cultural endemicity, is unique and
39 inherently valuable.

40

41 Indigenous Peoples have often constructed their ways of knowing using oral histories as one of the vehicles
42 of mind and memory, observance, governance, and maintenance of customary law (Table Cross-Chapter
43 Box INDIG.2). These ways of knowing can also incorporate the relationships between multiple factors
44 simultaneously which adds particular value towards understanding complex systems that is in contrast to the
45 dominant reductionist, Western approach- noting that non-reductionist approaches also exist (Ludwig et al.,
46 2014; Hoagland, 2017).

47

48 For climate research, the role of oral histories as a part of Indigenous knowledge and local knowledge is
49 extremely relevant. For example, ocean adaptation initiatives can be guided by oral historians and keepers of
50 knowledge who can convey new knowledge and baselines of ecosystem change over long-time frames
51 (Nunn and Reid, 2016). Oral histories can also convey cultural indicators and linguistic devices of species
52 identification as a part of a local dialect matrix and changes in ecosystems and species using interlinkages
53 not available to science (Mustonen, 2013; Frainer et al., 2020). Oral histories attached to maritime place
54 names, especially underwater areas (Brattland and Nilsen, 2011), can position observations relevant for
55 understanding climate change over long ecological timeframes (Nunn and Reid, 2016). Species abundances,
56 well-being and locations are some of the examples present in the ever-evolving oral histories as living ways
57 of knowing. Indigenous knowledge and oral histories may also have the potential to convey governance,

Do Not Cite, Quote or Distribute  18-76       Total pages: 197
    FINAL DRAFT                                         Chapter 18       IPCC WGII Sixth Assessment Report

1 moral, and ethical frameworks of sustainable livelihoods and cultures (Mustonen and Shadrin, 2020) rooted
2 in the particular Indigenous or local contexts that are not otherwise available in written or published forms.

3

4 Climate change research involving Indigenous Peoples and local communities has shown that the generation,
5 innovation, transmission, and preservation of Indigenous knowledge is threatened by climate change
6 (Kermoal and Altamirano-Jim閚ez, 2016; Simonee et al., 2021). This is because Indigenous knowledge is
7 taught, local knowledge is gained through experience, and relationships with the land are sustained through
8 social engagement within and among families, communities, and other societies (Tobias J.K, 2014; Kermoal
9 and Altamirano-Jim閚ez, 2016). The knowledge that has traditionally been passed on in support of identity,

10 language and purpose has been disrupted at an intergenerational level (Lemke and Delormier, 2017). Many

11 of these dynamics have affected local knowledge transfers equally (Mustonen, 2013). This scenario
12 represents a tension for Indigenous Peoples, where Indigenous knowledge in the form of land-based life
13 ways, languages, food security, intergenerational transmission and application are threatened by climate
14 change, yet in parallel, these same practices can enable adaptation and resilience (McGregor et al., 2020).

15

16

17 Table Cross-Chapter Box INDIG.1: Examples of Indigenous knowledge and local knowledge about climate change

18 used in this Assessment Report

    Issue           Examples of Indigenous Peoples' and local            Context, peoples, and Source
                                                                         location
                    communities' action

    Climate         Phenological cues to forecast and respond to climate Smallholder farmers,

    forecasting/ear change                                               Delta State, Nigeria

    ly warning      Forecasting of weather and climate variation         Afar pastoralists,

                    through observation of the natural environment (e.g. north-eastern Ethiopia Ch9

                    changes in insects, and wildlife).

    Fire hazards    Observation of wind patterns to plan response to     Inupiat, Alaska, US       Ch14
                    coastal erosion/flooding                                                       Ch12
                    Sky and moon observation to determine the onset of   Maya, Guatemala           Ch12
                    rainy season                                                                   Ch14
                                                                         Indigenous nations in
                    Prescribed burning                                   Venezuela, Brazil,        Ch12
                                                                         Guyana, Canada, and
    Crop yield /    Water management, native seeds conservation and      US
    food security   exchange, crop rotation, polyculture, and            Mapuche, Chile
                    agroforestry
    Livelihood      Crop association (milpa) agroforestry, land                                    Ch12
    and well-being  preparation and tillage practices, native seed
    Ecosystem       selection and exchange, adjusting planting           Maya, Guatemala
    degradation     calendars,
                    Harvesting rain-water and the use of maize           Yucat醤 Peninsula,        Ch14
                    landraces by Indigenous farmers to adapt to climate  Mexico
                    impacts and promote food security in Mexico
                    Cultural values ingrained in knowledge system:       Quechua, Cusco, Peru Ch12
                    reciprocity, collectiveness, equilibrium, and
                    solidarity                                           S醡i, Nenets, and         Ch13
                    Ecosystem restoration including rewilding            Komi, Scandinavia and     Ch14
                                                                         Siberia
                    Collaboration with researchers, foresters, and       Indigenous Nations in
                    landowners to manage native black ash deciduous      Canada and US
                    trees against emerald ash borer
                    Selection and planting of native plants that reduce
                    erosion

    Fisheries       Whole-of-island approaches that embed IK and LK      Small islands states (as  Ch15
                    in environmental governance                          defined by Chapter 15)
                    Traditional climate-resilient fishing approaches     Indigenous nations        Ch14
                                                                         across North America      CCP6
                                                                         and the Arctic

    Do Not Cite, Quote or Distribute                    18-77                                      Total pages: 197
   FINAL DRAFT                              Chapter 18           IPCC WGII Sixth Assessment Report

   Management Restoration of traditional network of water tanks  Traditional            Ch6

   of urban                                                      communities and

   resources                                                     activists in South

                                                                 Indian cities such as

                                                                 Bengaluru

1

2

3 Table Cross-Chapter Box INDIG.2: Case Study Summary

   Region       Summary

   Africa       Many rural smallholder farmers in Africa use their ingrained Indigenous knowledge systems to

                navigate climatic changes as many do not have access to Western systems of weather forecasting.

                Instead, these farmers have been reported to use observations of clouds and thunderstorms, and

                migration of local birds to determine the start of the wet season, as well as create temporary walls

                by rivers to store water during droughts. Indigenous knowledge systems should be incorporated

                into strategic plans for climate change adaptation policies to help smallholder farmers cope with

                climate change (Mapfumo et al., 2016).

   Arctic       For local Inuit hunters and others who travel across Arctic land, ice and sea, there is evidence that
                the most accurate approach to reduce risk and enable informed decision-making for safe travel, is
                to combine Indigenous knowledge and local observations of weather with official online weather
                and marine services information that is available nationally (Simonee et al., 2021). Combining
                Inuit and local knowledge of weather, water, ice, and climate information with official forecasts
                has provided local hunters with more accurate, locally relevant information, and has on several
                occasions helped to avoid major weather-related accidents.

   Latin        In Venezuela, Brazil, and Guyana, Indigenous knowledge systems have led to a lower incidence of
   America      wildfires, reducing the risk of rising temperatures and droughts (Mistry et al., 2016). The Mapuche

                Indigenous Peoples in Chile use various traditional and sustainable agricultural practices,
                including: native seed conservation and exchange (trafkintu), crop rotation, polyculture, and tree-

                crop association. They also give thanks to Mother Earth through rituals to nurture socioecological

                sustainability (Parraguez-Vergara et al., 2018). In rural Cusco Region of Peru, "cultures values
                known in Quechua as ayni (reciprocity), ayllu (collectiveness), yanantin (equilibrium) and
                chanincha (solidarity)" have led to successful adaptation to climate change (Walshe and

   Mori         Argumedo, 2016).
   (Aotearoa    The traditional calendar system (maramataka) used by the Mori in Aotearoa-New Zealand
   New          incorporates ecological, environmental and celestial Indigenous knowledge. Mori practitioners are
   Zealand)     collaborating with scientists through the Effect of Climate Change on Traditional Mori Calendars
                project (Harris et al., 2017) to examine if climatic changes are impacting the use of the
                maramataka, which can be used as a framework to identify and explain environmental changes.

                Observations are being documented across Aotearoa, New Zealand to improve understandings of

                environmental changes and explore the use of Indigenous Mori knowledge in climate change

   Skolt S醡i   assessment and adaptation.
   (Finland)    In 2011, the Skolt S醡i in Finland began the first co-governance initiative where collaborative
                management and Indigenous knowledge were utilized to effectively manage a river and Atlantic
                Salmon (Salmo salar). This species is culturally and spiritually significant to the Skolt S醡i and

                has been adversely impacted by rising water temperatures and habitat loss (Brattland and

                Mustonen, 2018; Feodoroff, 2020; Ogar et al., 2020) (see also CCP Polar). Using Indigenous

                knowledge, they mapped changes in catchment areas and used cultural indicators to determine the

                severity of changes. Through collaborative management efforts that utilized both Indigenous

                knowledge and science, spawning and juvenile habitat areas for trout and grayling were restored,

                demonstrating the autonomous community capacity (Huntington et al., 2017) of the Indigenous

                Skolt S醡i and the capacity of Indigenous knowledge to address climate change impacts and

                detection of very first microplastics pollution together with science (Pecl et al., 2017; Brattland and

                Mustonen, 2018; Mustonen and Feodoroff, 2020).

4

5 [END CROSS-CHAPTER BOX INDIG HERE]

6

7

8 18.4.3.2 Political and government arenas

 9

10 Climate resilient development is embedded in social systems, in the political economy and its
11 underlying ideologies, interests and institutions (see 18.4.1). The pursuit of CRD, and shifting

   Do Not Cite, Quote or Distribute                     18-78                           Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 development pathways away from prevailing trends, unfolds in an array of political arenas, from the
 2 offices of bureaucrats to parliament buildings, sidewalks and streets, to discursive arenas in which
 3 governance actors interact � from the village level to global forums (J鴕gensen et al., 2017; Montoute et
 4 al., 2019; S鴕ensen and Torfing, 2019; Pasquini, 2020). Paradoxically, the post-AR5 literature suggests that
 5 political arenas are often used to shut down efforts to explore the solution space for climate change and
 6 sustainable development (medium agreement, robust evidence) (e.g., Kenis and Mathijs, 2012; Kenis and
 7 Mathijs, 2014; Beveridge and Koch, 2016; Kenis and Lievens, 2016; Driver et al., 2018; Meriluoto, 2018;
 8 Swyngedouw, 2018; Mocca and Osborne, 2019). Power relationships among different actors create
 9 opportunities for people to be included or excluded in collective action (Sim閍nt-Germanos, 2019) (18.3.1.6,
10 18.4.3.5). Therefore, as evidenced by examples from the UK (MacGregor, 2019) and China (Huang and Sun,
11 2020) small-scale collective environmental action has transformative potential in part due to its ability to
12 increase levels of cooperation among different actors (medium agreement, limited evidence) (Green et al.,
13 2020; Bl黨dorn and Deflorian, 2021).

14

15 In addition to the `arm's length' acts of voting, social mobilisation, protest, and dissent can be critical
16 catalysts for transformative change (Porta, 2020). These are competitions for recognition, power, and
17 authority (Nightingale, 2017) that take place in settings. This is evidenced by experiences from the energy
18 sector in Bangladesh which became a contested national policy domain and where social movements
19 eventually transformed the nation's energy politics (Faruque, 2017). Similarly, in Germany, the nation's
20 energy transition led to marked changes in agency, legal frameworks, and energy markets drove the
21 proliferation of so-called municipalizations of energy systems � a reversal of years of system privatization
22 (Becker et al., 2016). Meanwhile, experience in Bolivia demonstrate that the transformative potential of
23 political conflict depends on transcending narrow issues to form broad coalitions with a collective identity
24 that challenge prevailing development objectives and trajectories (Andreucci, 2019). Such examples
25 illustrate the power of the communities as a vanguard against environmentally destructive practices
26 (Villamayor-Tomas and Garc韆-L髉ez, 2018). Social movements have been successful at countering fossil
27 fuel extraction (Piggot, 2018) and open up political opportunities in the face of increasing efforts to capture
28 natural resources (Tramel, 2018) and are bolstered by resistance from within some corporations and/or their
29 shareholders (Foug鑢e and Bond, 2016; Swaffield, 2017).

30

31 Coincident with these social movements targeting climate change and sustainability has been a rise of
32 political conservatism and populism as well as growth in misinformation (high agreement, medium evidence)
33 (Mahony and Hulme, 2016; Swyngedouw, 2019). This reflects efforts to maintain the status quo by actors in
34 positions of power in the face of rising social inertia for climate action (Brulle and Norgaard, 2019). Political
35 arenas of the future may even require a new body politic that includes non-humans and a new geo-spatial
36 politics (Latour et al., 2018).

37

38 As introduced in the discussion of governance as an enabling condition (18.4.2.1), a wide range of actors are
39 involved in successful adaptation, mitigation, and sustainability policy and practice including national,
40 regional and local governments, communities, and international agencies (Lwasa, 2015). As of 2018, 197
41 countries had between them over 1,500 laws and policies addressing climate change as compared to 60
42 countries with such legislation in 1997 when the Kyoto Protocol was agreed upon (Nachmany et al., 2017;
43 Nachmany and Setzer, 2018). In judicial branches, climate change litigation is increasingly becoming an
44 important influence on policy and corporate behavior among investors, activists, and local and state
45 governments (Setzer and Byrnes, 2019). There is enhanced action on climate change at both national and
46 subnational levels, even in cases where national policies are inimical as in USA (Carmin et al., 2012; Hansen
47 et al., 2013).

48

49 The strong role of governments in climate action has implications for the nature of democracy, the
50 relationship between the local and the national state, and between citizens and the state (Dodman and Mitlin,
51 2015). More integration of government policy and interventions across scales, accompanied by capacity
52 building to accelerate adaptation is needed (very high confidence). Key needs include enhanced funding,
53 clear roles and responsibilities, increased institutional capability, strategic approaches, community
54 engagement, judicial integrity (Lawrence et al., 2015). More resources, and more active involvement of the
55 private sector and civil society can help maintain adaptation on the policy agenda. Multilevel adaptation
56 approaches are also relevant in low-income countries where local governments have limited financial

Do Not Cite, Quote or Distribute  18-79       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 resources and human capabilities often leading to dependency on national governments and donor
 2 organizations (Donner et al., 2016; Adenle et al., 2017).

 3

 4 Unlike mitigation, adaptation has traditionally been viewed as a local process, involving local authorities,
 5 communities, and stakeholders (Preston et al., 2015). The literature on the governance of adaptation
 6 continues to emphasize that local governments have demonstrated leadership in implementation by
 7 collaborating with the private sector and academia. Local governments can also play a key role (Melica et
 8 al., 2018; Romero-Lankao et al., 2018) in converging mitigation and adaptation strategies, coordinating and
 9 develop effective local responses, enabling community engagement and more effective policies around
10 exposure and vulnerability reduction (Fudge et al., 2016). Local authorities are well-positioned to involve the
11 wider community in designing and implementing climate policies and adaptation implementation (Slee,
12 2015; Fudge et al., 2016). Local governments also help deliver basic services, and protect their integrity
13 from climate impacts (Austin et al., 2015; Cloutier et al., 2015; Nalau et al., 2015; Araos et al., 2017).
14 However, the resource limitations of local governments as well as their small geographic sphere of influence
15 suggests the need for more funding for this from higher levels of government, particularly national
16 governments, to address adaptation gaps (very high confidence) (Dekker, 2020). Local adaptation
17 implementation gaps can be linked to limited political commitment at higher levels of government and weak
18 cooperation between key stakeholders (Runhaar, 2018). Incongruities and conflicts can exist between
19 adaptation agendas pursued by national governments and the spontaneous adaptation practices of
20 communities. There may be grounds for re-evaluating current consultative processes integral to policy
21 development, if narrow technical approaches emerge as the norm for adaptation (Smucker et al., 2015).

22

23 Therefore, the traditional view of adaptation as a local process has now widened to recognize it as a multi-
24 actor process that transcends scales from the local and sub-national to national and even international (very
25 high confidence (Mimura et al., 2014). Many of the impacts of climate change are both local and
26 transboundary, so that local, bilateral and multilateral cooperation are needed (Nalau et al., 2015; Donner et
27 al., 2016; Magnan and Ribera, 2016; Tilleard and Ford, 2016; Lesnikowski et al., 2017). National policies
28 and transnational governance should be seen as complementary, especially where they favor transnational
29 engagement with sub- and non-state actors (Andonova et al., 2017). National governments typically act as a
30 pivot for adaptation coordination, planning, determining policy priorities, and distributing financial,
31 institutional and sometimes knowledge resources. National governments are also accountable to the
32 international community through international agreements. National governments have helped enhance
33 adaptive capacity through building awareness of climate impacts, encouraging economic growth, providing
34 incentives, establishing legislative frameworks conducive to adaptation, and communicating climate change
35 information (Berrang-Ford et al., 2014; Massey et al., 2014; Austin et al., 2015; Huitema et al., 2016).

36

37 18.4.3.3 Economic and financial arenas

38

39 The performance of local, national, and the global economies is a priority consideration shaping perceptions
40 of climate risk and the costs and benefits of different policy responses to climate change. The most
41 commonly used indicator of performance is gross domestic product (GDP) (Hoekstra et al., 2017).
42 Traditionally, national development efforts have sought to maximize the growth of GDP under the
43 assumption that GDP growth equates not only to economic prosperity (including poverty reduction) but also
44 to increased efficiency and reduced environmental externalities (Ota, 2017). Such assumptions often employ
45 models such as the environmental Kuznets curve (EKC) that postulates that economic development initially
46 increases environmental impacts, but these trends eventually reverse with continued economic growth.
47 Wealthy nations of the global North, including for example the United States, Great Britain, Iceland, Japan,
48 have had success over the past decade in reducing their greenhouse gas emissions while growing their
49 economies (very high confidence). However, attempts to empirically test EKC in different national contexts
50 has yielded mixed results. Case studies in Myanmar, China, and Singapore, for example, suggest that the
51 impacts of GDP on environmental quality are contingent on the development context and the environmental
52 impact under consideration (Aung et al., 2017; Lee and Thiel, 2017; Xu, 2018; Chen and Taylor, 2020). In
53 addition, an extensive literature now argues that current patterns of development, and the economic systems
54 underpinning that development, are unsustainable (Washington and Twomey, 2016), and thus economic
55 growth may not necessarily continue indefinitely in the absence of more concerted effort to pursue
56 sustainable development, including reducing the impacts of climate change.

57

Do Not Cite, Quote or Distribute  18-80       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 Given such criticisms of the link between development and economic growth, a growing number of
 2 researchers argue for the need for alternatives to GDP to guide development and evaluate the costs and
 3 benefits of different policy interventions (Hilmi et al., 2015). For example, while GDP growth can drive
 4 growth in income, it can also drive growth in inequality which can undermine poverty reduction efforts (very
 5 high confidence) (Fosu, 2017). Hence, recent years have seen significant interest in the concept of well-being
 6 as a more robust measure for linking policy and the economy with sustainable development for a healthy
 7 Anthropocene era (Fioramonti et al., 2019).

 8

 9 Another mechanism for evaluating environmental performance is to include environmental data in the
10 System of National Accounts (SNA) through the System of Environmental-Economic Accounting (SEEA)
11 introduced by the UN. As the international statistical standard for environmental-economic accounting
12 (Pirmana et al., 2019), SEEA includes natural capital resources in national accounting. A number of recent
13 studies conclude that failure to account for natural capital in macroeconomic impact assessments results in
14 overly optimistic outcomes (Pirmana et al., 2019; Jendrzejewski, 2020; Naspolini et al., 2020), (Banerjee et
15 al., 2019; Kabir and Salim, 2019; Keith et al., 2019). For example, Jendrzejewski (2020) inserted natural
16 capital into a computable general equilibrium model of the 2017 European windstorm on state-owned forests
17 in Poland. This resulted in more negative assessment of impacts, suggesting excluding natural capital could
18 lead to erroneous investments, strategies. or policies. Similarly, other studies rely on Quality of life (QOL)
19 measurements as alternatives for GDP. Estoque et al. (2018) suggested a "QOL-Climate" assessment
20 framework, designed to capture the social-ecological impacts of climate change and variability.

21

22 Another alternative to GDP is Green GDP which seeks to incorporate the environmental consequences of
23 economic growth (Boyd, 2007; Stjepanovi et al., 2017; Stjepanovi et al., 2019). Green GDP is difficult to
24 measure, because it is difficult to evaluate the environmental depletion and ecological damages of growth
25 (Stjepanovi et al., 2019). Although there is no consensus in measuring Green GDP, attempts have been
26 made for select countries including the United States (Garcia and You, 2017), Europe (Stjepanovi et al.,
27 2019), China (Chi and Rauch, 2010; Yu et al., 2019; Wang et al., 2020), Ukraine and Thailand
28 (Harnphatananusorn et al., 2019), and Malaysia (Vaghefi et al., 2015). Le (2016) illustrated the potential
29 negative impacts of climate change vulnerability on green growth. Some studies have suggested that
30 focusing on green growth as the only strategy to address climate change would be risky. Hickel and Kallis
31 (2020) argue that green growth is likely to be a misguided goal due to the difficulties of separating economic
32 growth from resource use and, therefore, carbon emissions (see also (Antal and van den Bergh, 2014).
33 Therefore, alternative strategies are required (Hickel and Kallis, 2020). In addition, green growth should also
34 be able to justly respond to social movements involving contestation, internal debates and tensions (Mathai
35 et al., 2018).

36

37 The emphasis on Green GDP is mirrored by another concept, Blue Growth, that focuses on the pursuing
38 sustainable development through the ecosystem services derived from ocean conservation (Mustafa et al.,
39 2019). Synthesis studies suggest that more intensive use of ocean resources, such as scaling up seaweed
40 aquaculture, can be used to enhance CO2eq sequestration, thereby contributing to greenhouse gas mitigation,
41 while also achieving other economic goals (Lilleb� et al., 2017; Froehlich et al., 2019). Similarly, Sarker et
42 al. (2018) present a framework for linking Blue Growth and climate resilient development in Bangladesh,
43 with Blue Growth representing an opportunity for adapting to climate change. Bethel et al. (2021) also links
44 Blue Growth to resilience, noting that a Blue economy can help facilitate recovery from the COVID-19
45 pandemic. Nevertheless, consistent with earlier assessment of enabling conditions for system transitions
46 (18.4.2.1), implementation of Blue Growth initiatives is contingent upon the successful achievement of
47 social innovation as well as creating an inclusive and cooperative governance structure (very high
48 confidence) (Larik et al., 2017; Soma et al., 2018).

49

50 A potential critique of the various alternative metrics and models for economic development is that they are
51 all framed in the context of growth. Over the past decade, ecological economists and political scientists have
52 proposed Degrowth (e.g., Kallis, 2011; Demaria et al., 2013) and managing without growth (e.g., Jackson,
53 2009) as a solution for achieving environmental sustainability and socio-economic progress. Such concepts
54 are a deliberate response to concerns about ecological limits to growth and the compatibility between
55 growth-oriented development and sustainability (Kallis et al., 2009). Sustainable degrowth is not the same as
56 negative GDP growth which is typically referred to as a recession (Kallis, 2011). Degrowth goes beyond
57 criticizing economic growth; it explores the intersection among environmental sustainability, social justice,

Do Not Cite, Quote or Distribute  18-81       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 and well-being (Demaria et al., 2013). Under current economic and fiscal policies (see Box 18.8), degrowth
 2 has been argued as an unstable development paradigm because declining consumer demand leads to rising
 3 unemployment, declining competitiveness, and a spiral of recession (Jackson, 2009: 46). More
 4 comprehensive modelling of socio-economic performance understands the segments of sufficient social
 5 transformation to guarantee maintenance and rise in wellbeing coupled with reduced 'footprints' (Raworth,
 6 2017; Hickel, 2019; D'Alessandro et al., 2020).

 7

 8

 9 START BOX 18.8 HERE

10

11 Box 18.8: Macroeconomic policies in support of Climate-Resilient Development

12

13 Climate change risk may differ from other economic and financial risks in a number of ways: climate change
14 is global; involves long-term impact; and involves a great deal of uncertainty; and with the possibility of
15 irreversible change (Hansen, 2021). The macroeconomic implications will differ across countries with less
16 developed countries are likely to suffer more relative to more advanced ones (Batten, 2018). Hence,
17 policymakers need to understand the impact of climate change on macroeconomic issues such as potential
18 output growth, capital formation, productivity, and long run level of interest rates, in order to better design
19 policy interventions, be it monetary or fiscal (Economides and Xepapadeas, 2018; Bank of England, 2019;
20 Rudebusch, 2019). As discussed, below a range of fiscal tools can be leveraged to mitigate the effects of
21 climate change (Krogstrup and Oman, 2019).

22

23 Monetary Policy

24

25 Changes in climate and subsequent policy responses could increase volatility of food and energy prices,
26 resulting in higher headline inflation rates. Thus, Central Banks (CBs) have to pay careful attention to
27 underlying inflationary factors in order to maintain their inflationary targets. In response, CBs can take a
28 number of actions. For example, they could require that collateral comprises assets that support the move to
29 low-carbon economy, or their refinancing operations and crisis facilities could incentivize borrowers' move to
30 low-carbon activities, particularly in countries where CBs' mandate has been expanded to account for climate
31 impact (Papoutsi et al., 2021). Other actions that CBs could take include adoption of sustainable and
32 responsible investment principles (Rudebusch, 2019), require financial firms to disclose their climate related
33 risks (ECB, 2020; Lee, 2020). Despite these opportunities, there is ongoing debate regarding whether CBs
34 should actively use monetary policy to address climate change and its risks (Honohan, 2019).

35

36 Fiscal policy

37

38 The application of green fiscal policies to address climate change could lead to environmental benefits
39 including environmental revenues that may be used for broader fiscal reforms (OECD, 2021). As the US aims
40 at becoming carbon neutral by 2050, fiscal policies at the national, sectoral, and international level can help to
41 achieve this goal, along with investment, regulatory, and technology policies (Parry, 2021). The effectiveness
42 of green fiscal policies are through their fiscal potential, opportunities for efficiency gains, distributional and
43 macroeconomic impacts, and their political economy implications (Metcalf, 2016). The International
44 Monetary Fund argues public support for green policies may rise in response to the COVID-19 crisis (IMF,
45 2017). For example, Leibenluft (2020) argues that investments to combat climate change should be an
46 important component of the efforts to rebuild the economy in the wake of COVID-19. Such action is justified
47 not only on ecological and social welfare grounds, but from a long-term fiscal perspective. For example,
48 climate change impacts and/or efforts to adapt to those impacts drive increased spending in areas such as public
49 health and disaster mitigation or response. Preventive and corrective actions would strengthen resilience to
50 shocks and alleviate the financial constraints they create, particularly for small countries (Catalano et al.,
51 2020). For example, Mallucci (2020) found that natural disasters exacerbate fiscal vulnerabilities and trigger
52 sovereign defaults in seven Caribbean countries. Ryota (2019) illustrates how to include natural disaster and
53 climate change in a fiscal policy framework to developing countries.

54

55 Carbon pricing

56

Do Not Cite, Quote or Distribute  18-82       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 Pricing of greenhouse gases, including carbon, is a crucial tool in any cost-effective climate change mitigation
 2 strategy, as it provides a mechanism for linking climate action to economic development (IMF/OECD, 2021).
 3 By 2019, 57 nations around the world had implemented or scheduled implementation of carbon pricing. These
 4 initiatives cover 11 gigatons of carbon dioxide or about 20% of greenhouse gases emissions. Carbon prices in
 5 existing initiatives range between $1 and $127 per ton of carbon dioxide, while 51% of the emissions that are
 6 covered are priced more than $10 per ton of carbon dioxide. Moreover, in 2018, Governments raised about
 7 $44 billion in carbon pricing revenues (World Bank, 2019). However, the carbon prices are lower than the
 8 levels required for attaining the ambitious goal of climate change mitigation, and therefore, prices would need
 9 to increase if pricing alone is going to be used to drive compliance with the Paris Agreement. Higher carbon
10 prices would also be warranted if prices are based on the social cost of carbon, which represents the present
11 value of the marginal damage to economic output caused by carbon emissions (Cai and Lontzek, 2018). This
12 cost needs to be considered with the social benefits of reducing carbon emissions through cost-benefit analyses
13 in order to make the intended regulation acceptable.

14

15 Taxes

16

17 Carbon taxes represent another financial mechanism for addressing climate (Metcalf, 2019), 2019b). For
18 example, the implementation of a carbon tax and a value-added tax on transport fuel in Sweden resulted in a
19 reduction of CO2 emissions from transport of about 11% in which the carbon tax had the largest share
20 (Andersson, 2019). In the United States, for example, a carbon tax could increase fiscal flexibility by collecting
21 new revenues that can be redeployed to finance reforms and help stimulate economic growth. However, U.S.
22 tax-inclusive energy prices would have to be 273% higher than laissez faire levels in 2055 in order to meet
23 international agreements (Casey, 2019). Similarly, limiting global warming to 2 degrees or less would likely
24 require a carbon tax rate in the Asia/Pacific region to be significantly higher than $25 per ton (IMF, 2021).
25 Therefore, using tax revenues to issue payments back to taxpayers that are disproportionately impacted or to
26 redistribute capital among regions may be one of the most important features of carbon tax policies. Although
27 the average effect of carbon tax on welfare would be positive, some regions (56%) will gain and some regions
28 (44%) lose (Scobie, 2013). Therefore, large transfer payments are needed to compensate those losing from
29 carbon tax (Krusell and Smith, 2018). IMF (2019) argues that, of the various mitigation strategies to reduce
30 fossil fuel CO2 emissions, carbon taxes are the most powerful and efficient, because they allow firms and
31 households to find the lowest-cost ways of reducing energy use and shifting toward cleaner alternatives.

32

33 Subsidies

34

35 The World Bank has been encouraging both developed and developing states, especially those with petroleum
36 reserves, to use the removal of subsidies as a mechanism for promoting energy transitions away from fossil
37 fuels. The transition has led to social unrest in some cases, especially where there is a culture of entitlement to
38 low-cost energy because it is an indigenous resource. Such reforms have been more effective when
39 governments have been able to clearly show how savings are applied to social and health programs that benefit
40 human well-being. Nevertheless, policy makers should not underestimate the complexity of issues involved in
41 the removal of subsidies that will increase the cost of carbon and hasten the transition to cleaner fuels (Scobie,
42 2017; Scobie et al., 2018; Chen et al., 2020a). A crucial issue to take into account is the harmful effects some
43 subsidies have on biodiversity. Although governments agreed in 2010 to make progress on reducing subsidies
44 in 2010, by 2020 few governments had identified specific incentives to remove or taken action toward their
45 removal. Further investigation of the positive and negative effects of subsidy redirection or elimination on
46 people and the environment (Dempsey et al., 2020).

47

48 END BOX 18.8 HERE

49

50

51 18.4.3.4 Knowledge-technology and ecological arenas

52

53 Knowledge-technology arenas comprise the interaction in knowledge spaces connected to technology
54 transitions. The institutional and political architecture through which knowledge and technology interact is
55 described in sustainability transitions literature (Fazey et al., 2018b; Sengers et al., 2019l Kanger, 2020
56 #3709). A common theme explored in that literature is the ability of actors to access and apply various forms
57 of knowledge as a means of effecting change. Different forms of innovation are recognized as a core

Do Not Cite, Quote or Distribute  18-83       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 enabling condition for achieving system transitions for CRD (18.3.3; Cross-Chapter Box INDIG). However,
 2 while scientific and technology knowledge may be useful, in some cases, they remain subordinate to political
 3 agendas, or are controlled by actors in positions of power and thus not equitably distributed (very high
 4 confidence) (Mormina, 2019). Participatory decision-making, for example, assumes that multiple actors,
 5 with differing motivations, agency and influence, engage with climate decision making and co-produce
 6 actions. Yet, some actors may not participate in the process if the proposed actions do not align with their
 7 motivations or if they do not have adequate agency (Roelich and Giesekam, 2019). Hence, effectively using
 8 knowledge to inform policy is challenging for both scientists, policymakers, and civil society alike.

 9

10 Science, technology, and innovation (STI) policies are expected to shape expectations of the potential for a
11 better world based on clean technologies, higher labor productivity, economic growth and a healthier
12 environment (Schot and Steinmueller, 2018; Mormina, 2019). STI policies are considered as `social goods
13 for development'. Hence, STI policies are often proposed or implemented as means of addressing
14 environmental challenges such as climate change along with sustainable development goals such as the
15 reduction of inequality, poverty, and environmental pollution (Mormina, 2019). Realizing the benefits of
16 STI, however, may be contingent on building broader STI capacity and bolstering nations' systems of
17 innovation (very high confidence) (Mormina, 2019). This could include building global research partnerships
18 to address priority STI needs as well as long-standing gaps between the global North and South. Such an
19 approach shifts the framing of STI as one focused on individual investigators to one comprised of building
20 knowledge networks. It also creates opportunities for integration of disparate forms of knowledge and
21 innovation, including local and indigenous knowledge, into global knowledge systems (Cross-Chapter Box
22 INDIG).

23

24 Furthermore, an extensive literature increasingly incorporates natural and ecological systems as knowledge
25 domains relevant to understanding opportunities for sustainability and CRD. For example, the literature on
26 socioecological systems (SES) (Sterk et al., 2017; Holzer et al., 2018; Avriel-Avni and Dick, 2019;
27 Mart韓ez-Fern醤dez et al., 2021) as well as social, ecological, and technological systems (SETS)
28 (McPhearson and Wijsman, 2017; Webb et al., 2018; Ahlborg et al., 2019), explicitly integrate ecological
29 knowledge into sustainability including concepts such as planetary boundaries (18.1.1), adaptation and
30 nature-based solutions, natural resources management, rights and access to nature, and understanding of how
31 humans govern society-nature interactions in the face of climate change (Benjaminsen and Kaarhus, 2018;
32 Mikulewicz, 2019; Nightingale et al., 2020). Some of these interactions are explained in Cross-Chapter Box
33 INDIG including conflict over which knowledges are recognized as valuable in understanding and
34 responding to climate change and therefore shape the nature of climate actions. Actor engagement in
35 stewardship, solidarity, inclusion of multiple knowledges and nature-society connectedness can highlight the
36 intertwined nature of ecological change and knowledge relations thereby support shifts to sustainability
37 (Pelling, 2010; Hulme, 2018; Ives et al., 2019; Nightingale et al., 2020) (see also Box 18.6).

38

39 The expanding definition of what constitutes credible, relevant, and legitimate knowledge is leading to the
40 democratization of knowledge and efforts to address historical inequities in access to knowledge (Ott and
41 Kiteme, 2016; Rowell and Feldman, 2019). This is reflected in the communication of science, which is
42 increasingly focused on reducing the distance between internal scientific and public communication and
43 more engagement in public science governance and knowledge production (Waldherr, 2012; Peters, 2013).
44 One innovative approach in co-production of knowledge is mobilizing communities through citizen science
45 (Heigl et al., 2019). This also presents additional opportunities to incorporate local knowledge with scientific
46 research, and better match scientific capability to societal needs.

47

48 18.4.3.5 Community arenas

49

50 Societal choices and development trajectories emerge from decisions made in different arenas which
51 intersect and interact across levels and scales, in diverse institutional settings - some formal with their
52 associated instruments and interventions, while others are informal. Since AR5, both formal and informal
53 setting are increasingly arenas of debate and contestation regarding development choices and pathways (very
54 high confidence) (see 18.4.4, Chapters 1, 6, 8, 10 and 17). Community arenas exist from the local to the
55 global scale and constitute the many interactions between governance actors, often transcending any one
56 scale to reflect the emergent outcomes of interactions in political, economic, socio-cultural, knowledge-
57 technology and ecological arenas of engagement. Actions within and between these five arenas hence come

Do Not Cite, Quote or Distribute  18-84       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 together in the community arena of engagement. While community engagement is often described at the
 2 level of villages and cities (Ziervogel et al., 2021) (Chapter 8), communities in terms of people interacting
 3 with each other sharing worldviews, values and behaviors, also exist at the regional and global levels. For
 4 example, civil society engagement in climate action reached a peak in 2019, notably through the global
 5 youth movement which led to large global mobilisation and street demonstrations on all continents and in
 6 many large cities (Bandura and Cherry, 2020; Han and Ahn, 2020; Martiskainen et al., 2020). Calling for
 7 enhanced climate action by governments and other societal actors, the youth movement was supported by
 8 many other societal groups and networks, including arenas of community interaction.

 9

10 While the SR1.5 (de Coninck et al., 2018) for the first time comprehensively assessed behavioral dimensions
11 of climate change adaptation, most literature still has a greater focus on what triggers mitigation behavior
12 (Lorenzoni and Whitmarsh, 2014; Clayton et al., 2015). Meanwhile, with CRD still a relatively young
13 concept, there is little literature focused on what motivates action in pursuit of CRD rather than its
14 subcomponents of climate action and sustainable development. Nevertheless, a common motivation that is
15 emerging in the literature is clinically significant levels of climate distress among individuals (Bodnar,
16 2008), which is experienced as a continuing distress over a changed landscape which no longer offers solace,
17 also known as solastalgia (high agreement, medium evidence) (Albrecht et al., 2007). This is accompanied
18 by a shift from blaming natural forces for disasters to attributing it to human negligence which is known to
19 lead to more acute perceptions of risk as well as more prolonged PTSD than trauma arising from non-human
20 causes. Improving social connections, acknowledging anxiety, reconnecting to nature, and finding creative
21 ways to re-engage are identified as ways of managing this growing anxiety (Lertzman, 2010; Clayton et al.,
22 2017). Climate action in communities at various scales could fulfil many of these needs.

23

24 18.4.4 Frontiers of Climate Action

25

26 After decades of limited government action and social inertia to reduce the risk of climate change, there is
27 also increasing social dissent toward the current political, economic and environmental policies to address
28 climate (Brulle and Norgaard, 2019; Carpenter et al., 2019). Social movements are demanding radical action
29 as the only option to achieve the mobilization necessary for deep societal transformation (very high
30 confidence) (Hallam, 2019; Berglund and Schmidt, 2020).

31

32 Prompted by SR1.5, new youth movements seek to use science-based policy to break with incremental
33 reforms and demand radical climate action beyond emissions reductions (Hallam, 2019; Klein, 2020;
34 Thackeray et al., 2020; Thew et al., 2020). Recent social movements and climate protests embrace new
35 modalities of action related to political responsibility for climate injustice through disruptive collective
36 political action (Young, 2003; Langlois, 2014). This is complemented by a regenerative culture and ethics of
37 care (Westwell and Bunting, 2020). These new social movements are based on nonviolent methods of
38 resistance, including actions classified as dutiful, disruptive and dangerous dissent (O'Brien, 2018).

39

40 The new climate movement mixes messages of fear and hope to propel urgency and the need to respond to a
41 climate emergency (Gills and Morgan, 2020). While some consider the mix between fear and hope as
42 beneficial to success depending on psychological factors (Salamon, 2019) or political geography (Kleres and
43 Wettergren, 2017) others warn of the risks of a rhetoric of emergency and its political outcomes (Hulme and
44 Apollo-University Of Cambridge Repository, 2019; Slaven and Heydon, 2020).

45

46 Research shows that new climate movements have increased public awareness, and also stimulated
47 unprecedented public engagement with climate change (very high confidence) (Lee et al., 2020; Thackeray et
48 al., 2020) and has helped rethink the role of science with society (Isgren et al., 2019). Such movements may
49 represent new approaches to accelerate social transformation and have resulted in notable political successes,
50 such as declarations of climate emergency at the national and local level, as well as in universities. Their
51 methods have also proven effective to end fossil fuel sponsorship (Piggot, 2018). Social demands for radical
52 action are likely to continue to grow, as there is growing discontent with political inertia and a rejection of
53 reformist positions.

54

55

56 [START BOX 18.9 HERE]

57

Do Not Cite, Quote or Distribute  18-85       Total pages: 197
    FINAL DRAFT                       Chapter 18               IPCC WGII Sixth Assessment Report

1 Box 18.9: The Role of the Private Sector in Climate Resilient Development via Climate Finance,

2   Investments and Innovation.

3

4 Climate finance broadly refers to resources that catalyze low-carbon and climate-resilient development. It
5 covers the costs and risks of climate action, supports an enabling environment and capacity for adaptation
6 and mitigation, and encourages R&D and deployment of new technologies. Climate finance can be
7 mobilized through a range of instruments from a variety of sources, international and domestic, public and
8 private (see Sections 18.4.2.2).

 9

10 The private sector has particular competencies which can make significant contributions to adaptation,
11 through innovative technology, design of resilient infrastructure, development and implementation of
12 improved information systems and the management of major projects. The private sector can be seen as a
13 "supplier of innovative goods and services" to meet the adaptation priorities of developing countries with
14 expertise in technology and service delivery (Biagini and Miller, 2013).

15

16 Future investment opportunities in CRD are in water resources, agriculture and environmental services.
17 Provision of clean water is another opportunity, requiring investment in water purification and treatment
18 technologies such as desalination, and wastewater treatment. Weather and climate services are a possible
19 area for private investment. (Hov et al., 2017; Hewitt et al., 2020).

20

21 [END BOX 18.9 HERE]

22

23

24 18.5 Sectoral and Regional Synthesis of Climate Resilient Development

25

26 Prior sections of this chapter assessed the literature relevant to CRD inclusive of climate risk management,
27 systems transitions and transformation, and actors and the arenas in which they engage one another to enable
28 or constrain CRD. Here, this knowledge is explored in different climatological and development contexts
29 through a synthesis of CRD-relevant assessments within the WGII sectoral and regional chapters.

30

31 18.5.1 Regional Synthesis of Climate-Resilient Development

32

33 In synthesizing regional knowledge relevant to the pursuit of CRD, this section first considers geographic
34 heterogeneity in regional responses of common climate variables to increases in globally averaged
35 temperatures. Such heterogeneity is a key driver of climate risk in different global regions, as well as human
36 and natural systems within those regions. This is followed by synthesis of various national development
37 indicators, aggregated to the regional level, as well as various challenges, opportunities, and options
38 supporting CRD reported within WGII regional chapters.

39

40 18.5.1.1 Climate Change Risk for Different Global Regions
41 Two important elements of understanding the opportunities and challenges associated with the pursuit of
42 CRD in different regional contexts are a) the geographic variability in climate conditions that shape
43 livelihoods, behaviors, and responses of human and natural systems; and b) how those conditions could shift
44 in the future in response to climate change, which determines the additional burden that climate change
45 could create for adaptation and sustainable development.

46

47 The climate analyses of WGI provide information on regional differences in temperature, rainfall, and sea-
48 surface temperatures for different global regions and how they are projected to change in response to
49 different levels of aggregate global warming (Table 18.4). Such data reveal that even when aggregated to
50 broad geographic regions, significant variations exist for all of these parameters, which is a function of the
51 baseline climatology of each region. For example, temperatures in Africa and Australia are, on average,
52 warmer than in Europe or North America. Significant variations are also observed for rainfall variables. Such
53 regional variation in climate conditions is part of the regional context that shapes current patterns of
54 development of the past present and future. They influence biodiversity and natural resource availability as
55 well as exposure to climatic extremes (tropical storms, heat waves, and drought) that contribute to disasters.

56

    Do Not Cite, Quote or Distribute  18-86                               Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 The WGI data also indicate that increases in globally averaged temperatures will have different
 2 consequences for regional climate change (Table 18.4), including variation in the magnitude and, for
 3 precipitation, even the direction of change (very high confidence). For example, although average
 4 temperatures, daily minimum temperature, and the number of days over a given thresholds are projected to
 5 increase in all regions except Antarctica, the magnitude of the change varies. Moreover, little change is
 6 projected for daily maximum temperatures across different regions. Nevertheless, the number of days over
 7 different temperature thresholds such as 35癈 increases markedly in most regions, reflecting the
 8 disproportionate impact that global warming has on the tails of temperature distributions. Given outcomes in
 9 many systems including public health, agriculture, ecosystems and biodiversity, and infrastructure are often
10 associated with biophysical thresholds (e.g., physiological or design thresholds), those regions where such
11 thresholds are increasingly exceeded due to climate change may experience disproportionately higher
12 impacts (very high confidence). Given such temperatures occur more frequently in regions such as Africa
13 and Central and South America, this disproportionate exposure is exacerbated by disproportionate
14 vulnerability, adaptation gaps, and development needs (very high confidence; 18.2.4; Table 18.4).

15

16 The regional response of precipitation to globally averaged temperatures increases is less clear than
17 temperature, in part due to high intra-region variability. Average daily precipitation remains fairly stable in
18 all global regions in response to higher magnitudes of global warming (Table 18.4). However, 5-day
19 precipitation totals provide a clearer signal of increasing hydrologic activity in response to higher globally
20 averaged temperatures (Table 18.4). Such data no not necessarily reflect changes in rainfall extremes that
21 could occur with downstream consequences for hazards such as drought or flooding. Similarly, while SSTs
22 are more uniform across global ocean basins, all basins are anticipated to warm in response to higher
23 globally averaged temperatures (Table 18.5). Unlike temperature, however, SST increases are anticipated to
24 be only a fraction of the globally averaged increase in temperature, due in large part to the heat capacity of
25 the oceans. Nevertheless, such higher SSTs have implications not only for ocean ecosystems and the
26 distribution of marine species, but also for weather patterns, such as formation and intensity of tropical
27 cyclones (very high confidence).

28

29 The other aspect of the regional climate responses to global temperature increases that is important for CRD
30 is the marked differences observed between changes in response to 1.5癈 versus 4癈 of warming. Higher
31 levels of global warming are associated with higher regional changes, including changes in extremes of
32 temperature. This in turn increases climate risk to exposed and vulnerable human and natural systems,
33 thereby increasing demand for adaptation. If that demand is not met, then the adaptation gap will be larger
34 with greater risk of loss and damage (very high confidence) (Schaeffer et al., 2015; Chen et al., 2016; United
35 Nations Environment Programme, 2021). This is true not only for regions, but also at the sectoral level
36 (18.5.2). Therefore, CRD pathways must balance the demands for emissions reductions to reduce exposure,
37 adaptation to manage residual climate change risks, and sustainable development to address vulnerability
38 and enhance capacity for sustainable development.

39

40

Do Not Cite, Quote or Distribute  18-87       Total pages: 197
   FINAL DRAFT                                    Chapter 18  IPCC WGII Sixth Assessment Report

1 Table 18.4: Projected continental level result ranges for select temperature and precipitation climate change variables by global warming level. Ranges are 5th and 95th percentiles
2 from SSP5-8.5 WGI CMIP6 ensemble results. There is little variation in the 5th and 95th percentile values by GWL across the SSP1-2.6, SSP2-4.5, SSP3-7.0, and SSP5-8.5
3 projections. Source: WGI AR6 Interactive Atlas (https://interactive-atlas.ipcc.ch/).

   Climate variable     Global       All Regions   North      Europe    Asia        Centra-South  Africa     Australia  Antarctica
                       warming                    America
                                                                                    America
                         level

                       4癈           12 to 15     8 to 11     5 to 9    12 to 14 24 to 27 26 to 29 24 to 27 -33 to -27

   Mean temperature 3癈              11 to 14     6 to 11     4 to 7    10 to 14    23 to 26      25 to 28   23 to 26 -35 to -26
                                     10 to 13     5 to 9      3 to 6     8 to 12    22 to 25      24 to 27   22 to 25 -36 to -27
   (degrees C)         2癈

                       1.5癈         9 to 12      4 to 8      2 to 5    8 to 12 22 to 24 24 to 26 22 to 24 -36 to -27

   Minimum of daily    4癈           -12 to -5 -25 to -15 -22 to -14 -18 to -9 11 to 15 10 to 14             5 to 10 -64 to -48

   minimum             3癈           -13 to -6 -27 to -15 -24 to -15 -20 to -11 10 to 15          8 to 14    4 to 10 -64 to -50

   temperatures         2癈          -15 to -8    -30 to -18 -27 to -17 -22 to -13  9 to 14       7 to 13    3 to 9     -65 to -51
    (degrees C)        1.5癈         -16 to -9    -32 to -20 -28 to -19 -23 to -14  8 to 14       6 to 12    3 to 9     -66 to -51

   Maximum of daily     4癈          32 to 37     32 to 38    28 to 33  35 to 40    36 to 43      40 to 47   41 to 49   -12 to -5
        maximum         3癈          31 to 39     31 to 38    28 to 34  35 to 41    35 to 44      39 to 51   41 to 54   -12 to -3
                        2癈          30 to 37     30 to 36    26 to 33  33 to 39    34 to 43      38 to 50   39 to 53   -13 to -4
      temperatures     1.5癈         29 to 36     29 to 35    25 to 31  32 to 39    33 to 42      38 to 49   39 to 52   -14 to -5
       (degrees C)

    Number of days      4癈          81 to 106    36 to 50    11 to 22  57 to 77    138 to 194 153 to 210 140 to 168    0 to 0
     with maximum       3癈          66 to 87     27 to 40     6 to 15  44 to 59    100 to 153 131 to 183 124 to 147    0 to 0
   temperature above    2癈          52 to 68     19 to 29     4 to 8   33 to 45     61 to 106 116 to 151 102 to 124    0 to 0
                       1.5癈         45 to 58     16 to 24     2 to 5   30 to 39     43 to 85 107 to 133 94 to 115      0 to 0
       35癈 � bias
         adjusted

   Near-surface total   4癈          2 to 3       2 to 3      2 to 2    2 to 3      4 to 5        2 to 3     1 to 2     1 to 1
      precipitation     3癈          2 to 3       2 to 3      2 to 2    2 to 3      3 to 5        2 to 3     1 to 2     1 to 1
        (mm/day)        2癈          2 to 3       2 to 3      2 to 2    2 to 3      3 to 5        2 to 3     1 to 2     1 to 1
                       1.5癈         2 to 3       2 to 3      2 to 2    2 to 3      3 to 5        2 to 3     1 to 2     1 to 1

   Maximum 5-day        4癈          79 to 99     75 to 93    53 to 71  81 to 105 118 to 168      68 to 113  81 to 124  20 to 29
     precipitation      3癈          66 to 99     68 to 87    48 to 68  70 to 101 97 to 165       60 to 118  76 to 129  19 to 27
    amount (mm)         2癈          64 to 93     65 to 84    47 to 65  66 to 95 93 to 162        55 to 107  73 to 122  18 to 26
                       1.5癈         63 to 91     63 to 83    46 to 64  64 to 93 92 to 160        52 to 105  74 to 119  18 to 25

4
5

   Do Not Cite, Quote or Distribute               18-88                       Total pages: 197
   FINAL DRAFT                       Chapter 18                                   IPCC WGII Sixth Assessment Report

1 Table 18.5: Projected sea surface temperature change ranges by global warming level and ocean biome (degrees Celsius). Ranges are 5th and 95th percentiles from SSP5-8.5 WGI
2 CMIP6 ensemble results. There is little variation in the 5th and 95th percentile values by GWL across the SSP1-2.6, SSP2-4.5, SSP3-7.0, and SSP5-8.5 projections. Source: WGI

3 AR6 Interactive Atlas (https://interactive-atlas.ipcc.ch/).

   Global   All ocean   Northern      Northern                 Equatorial   Southern    Southern   Gulf of           Eastern  Amazon   Arabian Indonesian
   warmin    biomes    Hemispher     Hemispher                             Hemispher   Hemispher   Mexico
   g level                                                                                                           Boundaries River  Sea      Flowthrough
                         e - High         e -                                   e -      e - High
                        Latitudes    Subtropics                            Subtropics   Latitudes

   4癈      1. t 2.    2. t 3.       2. t 2.                   2. t 3.     1. t 2.     1. t 2.     2. t 2.           2. t 2.  1. t 2.  2. t 2.  1.  t  2.7
            9 o 4      0 o 3         2 o 8                     1 o 0       8 o 4       3 o 0       1 o 8             1 o 7    7 o 5    3 o 9    9   o

   3癈      1. t 1.    1. t 2.       1. t 2.                   1. t 2.     1. t 1.     0. t 1.     1. t 2.           1. t 2.  1. t 2.  1. t 2.  1.  t  1.9
            3 o 7      2 o 2         4 o 4                     4 o 2       2 o 7       7 o 4       5 o 3             4 o 1    2 o 0    6 o 2    3   o

   2癈      0. t 1.    0. t 1.       0. t 1.                   0. t 1.     0.  t  1    0. t 0.     0. t 1.           0. t 1.  0. t 1.  0. t 1.  0.  t  1.2
            6 o 0      5 o 4         7 o 4                     7 o 3       5   o       3 o 8       6 o 4             6 o 3    6 o 3    6 o 3    5   o

   1.5癈    0. t 0. 0. t 0. 0. t 1.                            0. t 0.     0. t 0.     0. t 0. 0. t 1. 0. t 0.                0. t 0.  0. t 0.  0.  t  0.8
            2o7 1o9 2o0                                        2 o 8       2 o 6       1o5 2o0 2o9                            2 o 9    2 o 9    1   o

4

   Do Not Cite, Quote or Distribute                            18-89                               Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 18.5.1.2 Regional Perspectives on Climate-Resilient Development

 2

 3 The various regional chapters within the AR6 WGII report each provide insights into progress toward CRD
 4 as well as the opportunities and challenges associated with future pursuit of different CRD pathways.
 5 Common indicators of development reflect the significant diversity that exists across different global regions
 6 with respect to their development context (very high confidence). For example, the Human Development
 7 Index, recently adjusted to reflect the effect of planetary pressures (PPAHDI), illustrates the overall higher
 8 levels of development of North America and European countries of the global North as well as Australasia
 9 compared with Asia, Africa, Central and South America and small islands of the global South. Generally,
10 this reflects the higher levels of vulnerability and greater need for both sustainable development to reduce
11 poverty and support sustainable economies as well as climate action to address climate risk (Table 18.6).

12

13 However, even within a given region, there is significant variation in PPAHDI among nations. Such
14 differences reflect fundamental differences in historical patterns of development, as well as current
15 development needs and challenges, and they imply differences in what future development pathways would
16 be consistent with CRD. In addition, nations and regions with lower PPAHDI values suggests greater
17 capacity challenges for both greenhouse gas mitigation and climate adaptation. However, nations and regions
18 with high PPAHDI values also tend to have higher per capita CO2e emissions production, indicating that
19 economic development based on fossil fuel use undermines both efforts on climate action as well as the
20 SDGs (very high confidence) (Figure 18.6). Such challenges are also reflected by differential Gini
21 coefficients and metrics of state fragility among regions, which reflect inequities in income distribution and
22 broader vulnerability of nations and regions to shocks and stressors (Figure 18.6). In addition, high variation
23 is observed in CO2 emissions production, even among comparatively wealthy nations, suggesting CO2e
24 emissions of some nations are tightly coupled to development, while others have pursued more carbon
25 neutral development trajectories. Even within regions such as Africa, Asia, Central and South America, and
26 Europe, large within-region variations are observed in inequality and state fragility, suggesting high
27 variability among nations. Given the emphasis in the sustainable development and CRD literature on equity
28 and vulnerability, addressing such determinants of vulnerability is a core design principle for CRD
29 pathways.

30

31 In addition to development indicators, the literature assessed in the WGII regional chapters indicates that
32 different regions experience a range of development challenges and opportunities that affect the pursuit of
33 CRD (very high confidence). These represent dimensions of governance, institutions, economic
34 development, capacity, and social and cultural factors that shape decision-making, investment, and
35 development trajectories. For example, significant challenges exist within regions with respect to managing
36 debt and the ability to fund or finance climate action and sustainable development interventions (very high
37 confidence). On the other hand, a broad range of opportunities exist to pursue CRD including challenges
38 with debt and financing of adaptation competing policy objectives, social protection programs, economic
39 diversification, investing in education and human capital development, and expanding disaster risk reduction
40 efforts (very high confidence).

41

42

Do Not Cite, Quote or Distribute  18-90       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1

 2 Figure 18.6: Relationship among development indicators relevant to climate-resilient development. National Gini
 3 coefficients (most recent year available; n=141; (World Bank, 2021)), the Fragile States Index (2021; n=163; (Fund for
 4 Peace, 2021)), and per capita CO2 emissions (2018; n=169; (Human Development Report Office, 2020)) are plotted
 5 against the Planetary Pressures-Adjusted Human Development Index (2020, n=163; (Human Development Report
 6 Office, 2020)

 7

 8

 9 There are a wide variety of more focused options for climate action and sustainable development (very high
10 confidence). Such options have potential for synergies and trade-offs including implications for greenhouse
11 gas mitigation, land use change and conservation, food and water, or social equity. Despite variation in
12 development context, regional assessments suggest CRD efforts will be associated with some common
13 features. For example, in all regions, existing vulnerability and inequality exacerbate climate risk and
14 therefore pose challenges to CRD (very high confidence). Furthermore, low prioritization of sustainability
15 and climate action in government decision making, low perceptions of climate risk, and path dependence in
16 governance systems and decision-making processes all pose barriers to system transitions, transformation,
17 and CRD (very high confidence).

18

19 18.5.2 Sectoral Synthesis of Climate-Resilient Development

20 The sectoral chapters of the WGII report provide insights regarding how development processes interact
21 with sectors to shape the potential for climate-resilient development. Similar to global regions, each sector is
22 associated with various challenges, opportunities, and options that enable or constrain CRD (Table 18.7). A
23 number of challenges are common across sectors and mirror those associated with different regions. For
24 example, issues associated with natural resource dependency, access to information for decision-making,
25 access to human and financial capital, and path dependence of institutions represent barriers that must be
26 overcome if sectors are to support transitions that enable CRD. These challenges are more acute within
27 vulnerable communities or nations where capacity to innovate and invest are constrained and social
28 inequities reinforce the status quo (very high confidence). At the same time, a number of sector-specific
29 opportunities for mitigation, adaptation, and sustainable development can be used to integrate sectors into
30 CRD pathways. This could include policies and planning initiatives to enhance sector sustainability and
31 resilience as well as capacity building and greater inclusion of different actors and groups in decision making
32 including capitalizing on local and indigenous knowledge as a mechanism for more representative and
33 equitable action.

34 In addition, the sectoral assessments identify a broad range of specific adaptation, mitigation, and sustainable
35 development options that could play a role in facilitating CRD. Many of these options appear initially to be
36 specific to a given sector. For example, options for the water sector (Chapter 4) are assessed independently

Do Not Cite, Quote or Distribute  18-91       Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 from those for health and well-being (Chapter 7). In practice, however, evidence suggests the importance of
 2 thinking about sectoral options as cross-cutting, mutually supportive, and synergistic packages rather than
 3 singular options. First, each of the sectoral chapters has links to multiple SDGs (Table 18.7), implying each
 4 sector is important for achieving a range of sustainability goals that extend beyond sectoral boundaries.
 5 Moreover, progress across multiple sectors simultaneously creates opportunities for synergies for achieving
 6 the SDGs, but also enhances the risk of potential trade-offs (very high confidence). Second, a number of
 7 options are common to multiple sectors. For example, options associated with ecosystem-based adaptation
 8 and nature-based approaches to environmental management appear in multiple sectors (Table 18.7).
 9 Similarly, climate-smart agriculture and agroecological approaches to food systems create opportunities for
10 food security, but those same options also benefit land-based ecosystems, water, poverty and livelihoods,
11 and human well-being. Joint implementation

12

13 18.5.3 Feasibility and Efficacy of Options for Climate-Resilient Development

14

15 While both the sectoral and regional assessments indicate a rich toolkit of management options is available
16 to decision-makers to facilitate CRD, two key uncertainties undermine efforts to implement those options.
17 The first is the feasibility of implementation. Options that seem promising could nevertheless encounter
18 implementation barriers due to cost, absence of necessary capacity, lack of public acceptance, or competition
19 with alternative options. Progress in the literature since the AR5 and SR1.5 reports enables improved
20 consideration for options feasibility for both mitigation (SR1.5 ref) and adaptation (Cross-Chapter Box
21 FEASIB). This assessment allows the range of available options to be considered in a more critical light,
22 particular when on is considering opportunities for implementation over the near-term. Meanwhile, the other
23 challenge is that of option efficacy. Significant uncertainties remain regarding how well a given option will
24 perform in a specific context and whether it is capable of adequately addressing risk (18.6.1). Such
25 uncertainties can undermine the pursuit of CRD or at least efforts to accelerate system transitions that
26 support CRD (medium evidence, medium agreement) (18.3). Accordingly, closer examination of option
27 implementation in the real world, including within different sectoral and regional contexts, would enhance
28 the knowledge available to decision-makers regarding which options will best fit the needs of a given CRD
29 pathway.

30

31

Do Not Cite, Quote or Distribute  18-92       Total pages: 197
FINAL DRAFT                             Chapter 18                     IPCC WGII Sixth Assessment Report

1 Table 18.6: Regional synthesis of dimensions of climate-resilient development. For each region, quantitative information is provided on common development indicators including

2 the planetary pressures-adjusted human development index (PPHDI, 2020, n=169; (Human Development Report Office, 2020)), Gini coefficients (GINI, most recent year available;

3 n=156; (World Bank, 2021)), Fragile States Index (FRAGILITY; 2021; n=173; (Fund for Peace, 2021)), and per capita CO2 emissions production (CO2/PC, 2018; n=169; (Human

4 Development Report Office, 2020)). Each indicator is associated with a mean value among nations within a specific region as well as the range (minimum to maximum) value. In

5 addition, the table contains evidence of sustainable development challenges and opportunities as well as adaptation/sustainable development options and potential synergies and

6 trade-offs associated with their implementation. Synergies and trade-offs are categorized as follows: (T) Trade-off among policies and practices; (S+) Synergy among policies and

7 practices that enhances sustainability; (S-) Synergy among policies and practices that undermines sustainability.

Region       Development Indicators     Challenges                               Opportunities                       Options                             Synergies and Trade-Offs
                   mean (range)                                        � climate change literacy can

                                        � institutional and financial                                                � strengthening climate services � (T) competing uses for water

             PPAHDI      0.53           challenges in programming enable the mainstreaming of (9.4.2)                                                    such as hydropower
                        (0.39-          and implementing activities climate change into national � ecosystem based adaptation                            generation, irrigation, and

                        0.72)           to support concrete            and sub-national                                (9.11.4.2)                        ecosystem requirements
                        42.8            adaptation measures            developmental agendas                         � economic diversification          create trade-offs among

                 GINI   (27.6-            (9.14.5)                       (9.4.2)                                       (9.12.3)                             different management
             FRAGILITY  63.4)           � high debt levels exacerbate  � Adaptive responses can be                   � intensive irrigation9.15.2           objectives (9.7.3)
                         87.3                                                                                        � agricultural and livelihood       � (T) migration in response to
                        (57.0-            fiscal challenges and          used as an opportunity for                                                         unfavorable environmental
                        110.9)                                                                                         diversification (9.12.3)             conditions provides
                                          undermine economic             comprehensive,                              � drought resistant crop varieties     opportunities for farmers but
Africa                                                                                                                                                      puts pressure on the provision
                                          resilience (9.14)              transformative change                         (9.15.2)                             of social services and reduces
                                        � insufficient development       (9.6.2)                                     � soil and water conservation          farm labor (9.15.2)
                                                                       � Investments in human                                                            � (T) intensive Irrigation
                                          and adaptation finance and                                                   (9.15.2                              contributes to the
                                                                         capital, can facilitate
                                          accessibility of finance
                                                                         socioeconomic development
                                          (9.14.5)
                                        � complexity of estimating       and poverty reduction (9.9.1)
                                                                       � Strengthening the
                                          the costs and benefits for

                                        adaptation measures in         participation of women in                                                         development of agriculture

             CO2/PC               1.1   specific contexts (9.14.2)     decision-making as well as                                                        but has come at a cost to
                                                                       advance traditional and local                                                     ecosystem integrity and
                        (0.0-8.1) � exclusions of migrants and

                                        other vulnerable               knowledge can support                                                             human well-being (9.15.2)

                                        populations from social        climate action and

                                          programs (9.9.4)             sustainable livelihoods
                                        � mismatch between the         (9.9.3)

                                        supply of, and demand for,

                                        climate services (9.5)

                                  0.65 � migration and displacement � Investing in climate-resilient � risk insurance 10.5.5                             � (S+) nature-based adaptation

             HPAHDI     (0.47-          (Box 10.6)                     and sustainable                               � climate-smart agriculture         solutions, wetland protection,

                        0.78) � uneven economic                        infrastructure can be a                       10.4.5.5, (Table 10.6)              and climate-smart agriculture

Asia                              34.9  development (10.4.6)           source of green jobs as well � wetland protection and                             enhance carbon sequestration

             GINI       (26.6- � rapid land use change                 as a means of reducing                        restoration (Table 10.6)            (Table 10.6)

                        43.9)           (10.4.6)                       climate vulnerability

             FRAGILITY 73.6                                            (10.6.2)

Do Not Cite, Quote or Distribute                  18-93                                         Total pages: 197
FINAL DRAFT                         Chapter 18                     IPCC WGII Sixth Assessment Report

               CO2/PC     (32.3-    � increasing inequality        � sustainable development       � aquifer storage and recovery     � (S+) disaster risk reduction
                          111.7)      (10.4.6)                       pathways that connect                                              and capacity building has
               PPAHDI                                                climate change adaptation       (Table 10.6)                       synergistic interactions with
                 GINI       6.3     � large, socially                and disaster risk reduction   � integrated smart water grids       climate adaptation when the
                        (0.3-38.0)    differentiated vulnerable      efforts can reduce climate                                         two are effectively integrated
             FRAGILITY                populations (10.4.6)           vulnerability and increase      (Table 10.6)                       (10.6.2)
                           0.75                                      resilience (10.6.2)           � disaster risk management
               CO2/PC     (0.70-    � Underinvestment in                                                                              � (S+) environmental
               PPAHDI     0.81)       adaptation, particularly in  � social protection programs      (Table 10.6)                       sustainability has benefits for
                           34.4       public health systems,         can develop risk              � early warning systems (Table       relieving poverty and
                          (34.4-      given current and projected    management strategies to                                           promoting social equity
                          34.4)       risks (11.3.6.3)               address loss and damage         10.6)                              (10.6.4)
                           20.1                                      from climate change (10.5.6)  � resettlement and migration
                          (18.4-    � Underlying social and                                                                           � (T) intensive irrigation and
                          21.8)       economic vulnerabilities                                       (Table 10.6)                       other forms of water
                                      exacerbate disadvantage                                      � nature-based solutions in          consumption can have a
                           12.1       among particular social                                                                           negative effect on water
                        (7.3-16.9)    groups (11.8.2)                                                urban areas                        quality and aquatic
                                                                                                   � coastal green infrastructure       ecosystems (10.6.3)
                           0.71     � Competing policy and
                                      planning objectives within                                     (Table 10.6)                     � (T) adapting to fire risk in
                                      governments (11.7.2)                                                                               peri-urban zones introduces
                                                                   � implementation of national    � climate adaptation services,        potential trade-offs among
                                    � Limits to adaptation across    policies and guidance on                                            ecological values and fuel
                                      the region and among           climate adaptation and          planning and tools from             reduction in treed landscapes
                                      neighbors (11.7.2)             resilience (Box 11.5)                                               (11.3.5)
                                                                                                     government and private sector
                                    � Fear of litigation and       � cooperation among
                                      demands for compensation       individual farmers for          providers (11.7.1)
                                      create disincentives for       adaptation and regional       � enhancing governance
                                      climate adaptation (11.7.2)    innovation (11.7.1)
                                                                                                      frameworks (Table 11.17)
                                    � different climate change     � enhancing understanding of    � building capacity for
                                      risk perceptions among         Indigenous knowledge and
                                      different groups (11.7.2)      practices (Table 11.11)          adaptation (Table 11.17)
                                                                                                   � community partnership and
Australasia
                                                                                                      collaborative

                                                                                                      engagement(Table 11.17)
                                                                                                   � flexible decision-making

                                                                                                      (Table 11.17)
                                                                                                   � reducing systemic

                                                                                                      vulnerabilities (Table 11.17)
                                                                                                   � providing adaptation funding

                                                                                                      and compensation mechanisms

                                                                                                      (Table 11.17)
                                                                                                   � addressing social attitudes and

                                                                                                      engagement in adaptation and

                                                                                                      climate action (Table 11.17)

Do Not Cite, Quote or Distribute    18-94                          Total pages: 197
FINAL DRAFT                         Chapter 18                       IPCC WGII Sixth Assessment Report

Central          GINI    (0.62-     � vulnerability of informal      � Address existing               � upgrading of informal and         � (S+) conservation and
and South    FRAGILITY    0.78)                                        development deficits,                                                 restoration of natural
America                   47.2        settlements with chronic         particularly the needs of         vulnerable settlements              ecosystems have synergies
               CO2/PC    (38.6-                                        informal settlements and       � capacity building in national        with mitigation, adaptation
Europe                    57.9)       exposure to everyday, non-       economies                                                             and sustainable development
                          65.9                                                                           and city level government           (12.7.1)
                         (35.9-       climate risks                  � Adopt collaborative
                          92.6)     � limited political influence      approaches to decision-           institutions                     � (T) wind farms support
                                                                       making that integrate civic    � enhancing social protection          greenhouse gas mitigation but
                           2.2        of poor and most                 groups and communities as                                             have ecosystem implications
                        (0.9-4.8)                                      well as the private sector        programs                            and impacts (13.4.2)
                                      vulnerable groups                                               � integrated land use planning
                                    � poor market access of rural    � Enhance adoption of                                                � (T) adapting and mitigating
                                                                       sustainable tourism and           and risk-sensitive zoning           climate change through
                                      households                       livelihood diversification     � infrastructure greening              afforestation and forest
                                    � little consideration of the                                     � disaster risk mitigation and         management may be
                                                                                                                                             hampered by biophysical and
                                      implications of NDCs for                                           management                          land use trade-offs (13.3.2)
                                                                                                      � emergency medical and public
                                      poverty and livelihoods
                                    � corruption, particularly in                                        health preparedness
                                                                                                      � improving insurance
                                      the construction and
                                                                                                         mechanisms and climate
                                      infrastructure sector
                                    � gender inequities in labor                                         financing
                                                                                                      � ecosystem conservation,
                                      markets
                                    � limits to adaptation                                               protection, and restoration
                                                                                                      � appropriate use of climate
               PPAHDI      0.76     � mitigation and adaptation      � engagement in climate
                 GINI     (0.52-      remain siloed around             change knowledge, policy,         information and development
                          0.83)       sectoral approaches (Box         and practice networks (Box
             FRAGILITY     31.9       13.3)                            13.3)                             of climate services
                          (24.6-                                                                      � ecological restoration of
               CO2/PC     41.3)     � institutional, policy, and     � national policies can lead to
                           41.1       behavioral lock-ins              more ambitious and               habitats agroforestry and
                          (16.2-      constrain the rate of system     integrated climate planning
                          72.9)       transitions (13.11.4)            and action with associated       reforestation (13.8.2)
                                                                       co-benefits (Box 13.3)         � "smart farming" and
                            6.8     � legislative and decision-
                        (1.3-21.3)    making process constraints     � system transformations           knowledge training (13.5.2.1)
                                      on climate action (13.11.4)      towards more adaptive and      � soil management practices
                                                                       climate resilient systems
                                    � high adaptation costs and        (13.11.4, Box 13.3)              (13.5.2.1)
                                      concerns about                                                  � changing sowing dates and
                                      effectiveness and feasibility
                                      (13.3.2, Table 13.A.5)                                            changes in cultivars (13.5.2.1)
                                                                                                      � stricter enforcement of existing
                                    � competition for land use
                                      among adaptation and other                                        health regulations (13.7.2)
                                      uses (13.3.2)                                                   � integrated coastal zone

                                                                                                        management and marine

                                                                                                        spatial planning (13.4.2)
                                                                                                      � nature-based solutions (13.4.2)
                                                                                                      � climate services 13.6.2.3

Do Not Cite, Quote or Distribute    18-95                            Total pages: 197
FINAL DRAFT                         Chapter 18                       IPCC WGII Sixth Assessment Report

North          PPAHDI      0.72     � perceptions of climate         � increased focus on building   � tailored insurance products for  � (S+) Post-fire ecosystem
America          GINI     (0.72-       change as irrelevant or not     adaptive capacity in small                                          recovery measures,
                          0.73)        urgent (13.3.2)                 towns and rural areas           specific physical climate           restoration of habitat
Small        FRAGILITY     40.0                                        (14.6.3)                                                            connectivity, and managing
Islands                   (33.3-    � public budget and human                                          risks13.6.2.5                       for carbon storage enhance
               CO2/PC     45.4)        capital limitations (13.3.2)  � greater use of SDGs as a      � protection of world heritage        adaptation potential and
                           45.4                                        framework for equitable                                             offers co-benefits with carbon
               PPAHDI     (21.7-    � lack of representation of all    adaptation measures (14.6.3)    sites (13.8.2)                      mitigation (Box 14.1)
                 GINI     69.9)        groups and communities in                                     � indigenous knowledge-based
                                       politics and decision-        � broader and deeper                                               � (T) REDD+ represents a
             FRAGILITY     11.9        making (14.6.3)                 recognition of the role of       land and resource management       trade-off between carbon
                        (3.8-16.6)                                     Indigenous knowledge and                                            mitigation and the ability of
                                    � economic and financial           local knowledge systems in       (Section 14.4.4)                   communities to improve their
                           0.68        constraints on adaptation       adaptation (14.6.3)           � adaptive co-management of           food security (14.4.7)
                          (0.51-       within communities 14.6.2
                          0.76)                                      � greater emphasis on              agriculture and freshwater      � (T) New coastal and alpine
                           40.2     � persistent social                participatory governance and                                        developments generate
                          (28.7-       vulnerability and inequities    co-production of knowledge       resources (Section 14.4.3)         economic activity but
                          56.3)        14.6.3, 14.4.7.3                in adaptation decision-       � ecosystem based management          enhance local social
                           64.6                                        making (14.6.2.2)                                                   inequalities (15.4.10)
                          (38.1-    � adaptation actions that are                                       and nature based solutions
                          97.5)        maladaptive and exacerbate    � enhanced use of risk-based                                       � (S+) development decisions
                                       existing inequities             decision analysis                (Box 14.3) Section 14.4.2,         and outcomes are
                                       (14.6.2.1)                      frameworks and flexible                                             strengthened by consideration
                                                                       adaptation pathways              14.4.3, 14.4.4) (Table 14.9).      of climate and disaster risk
                                    � constraints on capacity for      (14.6.2.2)                    � increase efficiency and equity      (15.7)
                                       data collection (Table 14.8)
                                                                     � coordination of policies to      of water management and         � (S-) impacts of invasive alien
                                    � limited organizational           support transformational                                            species on islands are
                                       willingness implement new       adaptation (14.6.2.2)            allocation (14.4.3.3)              projected to increase with
                                       and untested solutions                                        � energy conservation measures
                                       (Table 14.8)                  � increasing women's access
                                                                       to climate change funding        (14.6.1.3)
                                    � high dependence of               and support from              � guidelines, codes, standards,
                                       economic activity on            organizations
                                       tourism (15.3.4.5)              (15.6.5)promoting                and specifications for
                                                                       agroecology, food
                                    � Lack of coordination             sovereignty, and                 infrastructure (14.6.1.6)
                                       among government                regenerative economies        � modifying zoning and buying
                                       departments (15.6.1)            (15.7)
                                                                                                        properties in floodplains
                                    � limited regional
                                       cooperation (15.6.1)                                             (14.6.1.3)
                                                                                                     � web-based tools for

                                                                                                        visualizing and exploring

                                                                                                        climate information scenario

                                                                                                        planning and risk analyses
                                                                                                        (s14.6.1.6)
                                                                                                      � raising dwellings and other

                                                                                                         infrastructure (15.5.2)
                                                                                                      � land reclamation (15.5.2)
                                                                                                      � migration and planned

                                                                                                         resettlement (15.5.2)
                                                                                                      � ecosystem-based adaptation

                                                                                                         including Indigenous and

                                                                                                         local knowledge (15.5.2)

Do Not Cite, Quote or Distribute    18-96                            Total pages: 197
   FINAL DRAFT                             Chapter 18                      IPCC WGII Sixth Assessment Report

                                           � absence of planning           � expanding sustainable         � protected areas (15.5.2)           time due to synergies between
                                                                             tourism economies (15.7)      � ecosystem restoration and          climate change and other
                                              frameworks (15.6.1)                                                                               drivers (15.3.3)
                                           � corruption and corrupt        � integrating climate change       improved agroforestry          � (S-) synergies between
                                                                             and disaster management                                            changing climate and other
                                              people in political and        with broader development         practices (15.5.2 15.5.4)         natural and anthropogenic
                                                                             planning and                  � community-based adaptation         stressors could lead to
                                              public life (15.6.1)           implementation (15.7)                                              disproportionate impacts on
                                           � insufficient human capital                                       (15.5.5)                          biodiversity (15.3.3)
                                                                           � using climate risk insurance  � livelihood diversification and
                     CO2/PC        3.7        (15.6.1)                       as a way to support
                               (0.3-31.3)  � competing development           development and adaptation       use of improved technologies
                                                                             processes (15.7)
                                              priorities (15.5.5)                                             and equipment (15.5.6)
                                           � lack of education and         � improving cross sectoral and  � diversifying cropping
                                                                             cross agency coordination
                                              awareness around climate       (15.7)                           patterns, expanding or

                                              change (15.6.4)              � enhanced integration             prioritizing other cash crops
                                           � failure of externally driven    between development
                                                                             assistance, public financial     (15.5.6)
                                              adaptation (15.6.5)            management, and climate       � small-scale livestock
                                           � constraints on economic,        finance (15.5.7)
                                                                                                              husbandry (15.5.6)
                                              legislative, and technical                                   � irrigation technologies

                                              capacity of local                                               (15.5.6)
                                                                                                           � diversification away from
                                              governments (15.7)
                                                                                                              coastal tourism
                                                                                                           � disaster risk management

                                                                                                              (DRM) (15.5.7)
                                                                                                           � early warning systems and

                                                                                                           climate services (15.5.7)

1

2

3 Table 18.7: Sectoral synthesis of dimensions of climate-resilient development. For each sectoral chapter of the WGII report, this table identifies those SDGs that are discussed in the

4 relevant chapter as being particularly relevant to the sector. In addition, the table contains evidence of sustainable development challenges and opportunities as well as

5 adaptation/sustainable development options and potential synergies and trade-offs associated with their implementation. Synergies and trade-offs are categorized as follows: (T)

6 Trade-off among policies and practices; (S+) Synergy among policies and practices that enhances sustainability; (S-) Synergy among policies and practices that undermines

7 sustainability.    Relevant              Challenges                        Opportunities                         Options                             Trade-offs
             Sector   SDGs                                       � nature based solutions offer the  � habitat restoration,             � (S+) ecosystem-based adaptation

                     SDG 1, � low capacity for dispersal limits

   Terrestrial       SDG 2, range shifts to match climate        opportunity to address climate            connectivity, and creation of measures, such as restoration of

   and               SDG 3, (2.6.1)                              change and biodiversity               protected areas (Table 2.5)           forests and wetlands for flood and
   freshwater        SDG 6, � constraints on the evolution of    problems in an integrated way       � integrated landscape                  erosion control help maintain

   ecosystems        SDG 7, greater stress tolerance among         (2.6)                                   management (Table Cross-          freshwater supply and quality
   and their                                                     � adaptation can be integrated
   services          SDG 9, species (2.4.2, 2.6.1)                                                         Chapter Box NATURAL.1 in (2.2.2)
                     SDG 10, � altered peatland drainage and       with the protection of
                                                                                                           Chapter 2)                   � (S-) over-grazing/stocking of

                     SDG 11, repeated disturbances pose          biodiversity and land-based                                                 pastures and grasslands can result

   Do Not Cite, Quote or Distribute                    18-97                                     Total pages: 197
FINAL DRAFT                       Chapter 18                 IPCC WGII Sixth Assessment Report

             SDG 12,    barriers to restoration of tropical     climate change mitigation       � community-based natural            in soil erosion and the loss of
             SDG 13,    peatlands (2.4.3)                       initiatives (2.6.2)                                                  biodiversity (Table Cross-Chapter
             SDG 15,  � demonstrating the efficacy of                                              resource management               BoxNATURAL1 in Chapter 2)
             SDG 17     natural flood management efforts     � development assistance can help                                    � (T) planting non-native
                        poses challenges to its                address resource constraints        (2.6.5.7)                         monocultures for mitigation can
                        deployment (2.6.5)                     associated with marine           � maintain or restore natural        reduce biodiversity and resilience
                      � uncertainties in climate and           ecosystem management (3.6.3)                                       � (T) inappropriate hydrological
                        socioeconomic projections                                                  species and structural            restoration can result in increased
                        constrain adaptation planning and    � improving coordination among                                          methane emissions (Table Cross-
                        implementation (2.7)                   actors and projects will            diversity (Table Cross-           Chapter Box NATURAL1 in
                                                               contribute to achieving SDGs                                          Chapter 2)
Ocean and    SDG1,    � shifts in the distribution of fish     (3.6.3)                             Chapter Box NATURAL.1 in       � (T) afforestation/reforestation and
coastal      SDG2,      species across exclusive                                                                                     bioenergy initiatives can conflict
ecosystems   SDG3,      economic zones present               � private finance can support         Chapter 2)                        with other land uses such as food
and their    SDG5,      governance, ecological, and            restoration of blue-carbon       � restoration of hydrological        and timber production (Table
services     SDG7,      conservation challenges (3.4.3)        systems (3.6.3)                                                       Cross-Chapter Box BECCS, 2.2.2,
             SDG8,                                                                                 flows and catchment               Box 2.2)
             SDG9,    � resource constraints impede the      � joint implementation of coastal
             SDG10,     implementation of ecosystem-           and marine management               vegetation (Table Cross-       � (S+) adaptation in ocean and
             SDG11,     based and community-based              initiatives can address                                              coastal systems can be designed in
             SDG12,     adaptation for low- to middle-         governance challenges across        Chapter Box NATURAL.1 in         ways that substantially contribute
             SDG13,     income nations (3.6.2)                 scales and sectors (3.6.3)                                           to the SDGs and not only support
             SDG14                                                                                 Chapter 2)                       but allow the attainment of social,
                      � governance in marine social-         � ocean-based renewable energy     � control of feral herbivores       environmental and economic
                        ecological systems is highly           options can reduce reliance on                                       targets (3.6.4)
                        complex with poorly-defined            imported fuel (3.6.3)               with (Table Cross-Chapter
                        legal frameworks (3.6.2)                                                                                  � (S+) blue/green economies can
                                                                                                   Box NATURAL.1 in Chapter         reduce emissions and finance
                      � "Coastal squeeze" challenges                                                                                adaptation pathways (3.6.3)
                        adaptation, creating tensions                                              2)
                        between coastal development and                                         � reduce non-climatic stressors   � (T) built infrastructure conflicts
                        coastal habitat management                                                                                  with mitigation goals and can
                        (3.6.3)                                                                    to land-based ecosystems         create potential ecological, social
                                                                                                                                    and cultural impacts that
                                                                                                   (Table 2.6)                      undermines ecosystem health
                                                                                                � maritime spatial planning and     (3.6.2)

                                                                                                  integrated coastal

                                                                                                  management (3.6.2; Figure

                                                                                                  3.2.6)
                                                                                                � adaptive and sustainable

                                                                                                  fisheries management (3.6.2)
                                                                                                � habitat restoration (3.6.2)
                                                                                                � fishery mobility (Figure

                                                                                                  3.6.2)
                                                                                                � assisted evolution (Figure

                                                                                                  3.2.6)
                                                                                                � increase participation in

                                                                                                  management and governance

                                                                                                  (Figure 3.2.6)
                                                                                                � nature-based solutions (3.6.2)
                                                                                                � hard and soft infrastructure

                                                                                                  (Figure 3.2.6)
                                                                                                � livelihood diversification

                                                                                                  (Figure 3.6.2)

Do Not Cite, Quote or Distribute  18-98                      Total pages: 197
FINAL DRAFT                       Chapter 18                 IPCC WGII Sixth Assessment Report

Water         SDG 1,   � uncertainty in future water         � a resilient circular economy       � disaster mitigation and          � (S+) increasing the proportion of
              SDG 2,     availability (Box 4.1, Box 4.4)       delivers access to water,                                                sewerage, treated wastewater,
Food, fiber,  SDG 3,                                           sanitation, wastewater, and          response (Figure 3.2.6)             recycling and safe reuse would
and other     SDG 6,   � lack of sufficient data,              ecological flows (Box 4.7)         � finance and market                  help reach climate and water
ecosystem     SDG 7,     information and knowledge in                                                                                   targets (Box 4.7)
products      SDG 10,    understanding the water energy      � adaptive sanitation systems and      mechanisms (Figure 3.2.6)
              SDG 11,    food nexus (Box 4.6)                  sustainable urban drainage         � changes in crop cultivars and    � (S+) solar irrigation pumps provide
              SDG 13                                           contribute to a `one health                                              for income diversification for
                       � increasing urbanization is            approach' which can prevent          agronomic practices (4.5)           small and marginal farmers while
               SDG1,     creating new and difficult            water and sanitation               � changes in irrigation and           also generating renewable energy
               SDG2,     demands for urban water               contamination risks during                                               (Box 4.7)
               SDG3,     management. (4.3.4)                   floods and droughts. (Box 4.7)       water management practices
               SDG4,                                                                                                                 � (T) desalination of seawater or
               SDG5,   � barriers to adapting water-         � climate-proof infrastructure         (4.5)                               brackish inland water is energy-
               SDG6,     dependent livelihoods in rural        would reduce infection risks in    � water and soil conservation         intensive, high salinity brine, and
                         communities (4.3.1)                   flood-prone areas (Box 4.7)                                              other contaminants (4.5.5)
                                                                                                    (4.5)
                       � mainstreaming water                 � governance can derive              � migration and off-farm           � (T) negative-emission
                         management across sectors and         legitimacy from inclusion of                                             technologies, such as direct air
                         enhancing finance for adaptation      multiple stakeholders, including     livelihood diversification          capture can result in a net increase
                         (4.3.5)                               women, indigenous communities                                            in water consumption (4.5.5)
                                                               and young people (4.6.6)             (4.5)
                       � path-dependency of institutions,                                         � collective action, policies and  � (S+) agricultural production
                         and the speed at which these        � Indigenous and local knowledge                                          systems that integrate crops,
                         allow for changes in the decision-    can help ensure solutions align      institutions (4.5)                 livestock, forestry, fisheries and
                         making process (4.5.3)                with the interests of communities  � economic and financial             aquaculture can increase food
                                                               (FAQ 4.5)                                                               production per unit of land, reduce
                       � increased cost and management                                              incentives (4.5)
                         challenges of providing safe food   � integrated approaches to food,     � training and capacity building
                         (5.2.2)                                water, health, biodiversity and
                                                                energy that involve vulnerable      (4.5)
                       � warming-induced shifts of species      groups can help to address        � flood risk reduction measures
                         create resource allocation             current and future food security
                                                                challenges, reduce vulnerability    (4.5)
                                                                                                  � urban water management

                                                                                                    (4.5)
                                                                                                  � water, sanitation, and hygiene

                                                                                                    adaptations (4.5)
                                                                                                  � agro-forestry and forestry

                                                                                                    responses (4.5)
                                                                                                  � livestock and fishery

                                                                                                    responses (4.5)
                                                                                                  � indigenous and local

                                                                                                    knowledge (4.5)
                                                                                                  � energy related adaptations

                                                                                                    (4.5)
                                                                                                  � livelihood diversification

                                                                                                    (5.4.4)
                                                                                                  � social protection policies and

                                                                                                    programs (5.4.4)
                                                                                                  � changes in crop management

                                                                                                    including irrigation,

Do Not Cite, Quote or Distribute  18-99                      Total pages: 197
FINAL DRAFT                       Chapter 18                   IPCC WGII Sixth Assessment Report

                SDG7,     challenges among different              of Indigenous people, small-         fertilizers, planting schedules,    climatic risk, and reduce emissions
                SDG9,     fishing fleets (5.2.1)                  scale landholders and                                                    (Chapter 5 ES)
                SDG9,   � challenges related to REDD+             pastoralists, and promote            and crop varieties (5.4.4.1)      � (S+) integrated approaches to food,
                SDG10,    implementation and forest use           resilient ecosystems. (5.12.3,     � adjusting water management          water, health, biodiversity and
                SDG11,    (5.6.3)                                 5.13.2; 5.14)                                                            energy can help address current
                SDG12,  � differences in perceptions about     � agroforestry delivers benefits for    for forage production (5.5.4)       and future food security
                SDG13,    the validity of different forms of      climate change mitigation,         � rotational grazing of livestock     challenges, reduce vulnerability of
                SDG14,    knowledge (5.8.4)                       adaptation, desertification, land                                        Indigenous people, small-scale
                SDG15,  � inequality in access to climate         degradation, and food security       (5.5.4)                             landholders and pastoralists, and
                SDG16     services (5.14.1)                       and is considered to have broad    � fire management to control          promote resilient ecosystems.
                        � lack of support, policies, and          adaptation and moderate                                                  (5.12.3, 5.13.2; 5.14)
                SDG11,    incentives for the adoption of          mitigation potential (5.10.4)        woody thickening of grass         � (T) growing biomass demand for
                SDG13,    agroecological approaches            � partnerships between key                                                  producing sustainable bioproducts
                SDG17     (BIOECO.1)                              stakeholders such as researchers,    (5.5.4)                             competes with food production
                        � financial barriers limit                forest managers, and local actors  � using more suitable livestock       with potential effects on food
                          implementation of adaptation            can lead to a shared                                                     prices and knock-on effects related
                          options in agriculture, fisheries,      understanding of climate-related     breeds or species (5.5.4)           to civil unrest (BIOECO.1)
                          aquaculture and forestry (5.14.3)       challenges and more effective      � migratory pastoralist
                                                                  decisions. (5.6.3)
                         � poor municipal funding, data        � urban ecological infrastructure       activities (5.5.4)
                            collection, and collaboration         including green, blue, turquoise   � monitor and manage the
                            hinders sustainable development       and others can be a source of
                            initiatives, capacity building,       nature-based solutions that can      spread of pests, weeds, and
                            and climate action (6.1.5, 6.4.5,     improve both adaptation and
                            6.4.9)                                mitigation in urban areas (6.1.2)    diseases (5.5.4)
                                                               � transition architecture             � nature- or ecosystem-based
                         � high urbanization rates pose           movements can drive urban
                            challenges to areas that already      adaptation (6.4.1)                   strategies (5.12.5.2)
                            have high levels of poverty,       � transformative capacities
Cities,                     unemployment, informality, and        support adaptation efforts and     � green infrastructure,             � (S+) sustainable urban energy
settlements                 housing and service backlogs          systemic change processes                                                planning that includes
and key                     (6.2.1)                               (6.4.4)                              sustainable land use and            opportunities to avoid and reduce
infrastructure                                                 � incorporating Indigenous and                                              the UHI effect can provide
                         � Limited capacity for early-            local knowledge help generate        planning, and sustainable           synergies for both climate
                            warning systems in low-income         more people-oriented and place-                                          mitigation and adaptation in urban
                            countries (6.3.2)                     specific adaptation policies         water management (6.1.2)            areas (Cross-Chapter Box URBAN
                                                                  (6.4.7)                            � nature-based solutions (6.3.3)      in Chapter 6)
                         � lack of administrative capacities,  � climate finance offers the          � insurance (6.3.2)
                            coordination across sectors and       opportunity to overcome            � switching to air cooling for      � (S+) natural ventilation and
                            efforts, transparency and             structural impediments to                                                passive energy strategies can
                            accountability slows                  climate action (Box 6.5)             thermal power plants (6.3.4)        capture synergies between climate
                            sustainability transitions and                                           � increasing the efficiency of        mitigation and adaptation (Cross-
                            disaster risk reduction (Case                                                                                  Chapter Box URBAN in Chapter
                            Study 6.4)                                                                 hydro and thermoelectric            6)

                                                                                                       power plants (6.3.4)              � (S+) community-based adaptation
                                                                                                     � changing reservoir operation        has potential to be better integrated
                                                                                                                                           to enhance well-being and create
                                                                                                       rules (6.3.4)                       synergies with the Sustainable
                                                                                                     � upgrading infrastructure and        Development Goals

                                                                                                       strengthening, or relocating      � (T) urban mitigation efforts can
                                                                                                                                           create trade-offs with adaptation
                                                                                                       (critical) assets (6.3.4)           such as intensifying the Urban
                                                                                                     � including green, blue,

                                                                                                       turquoise and nature-based

                                                                                                       solutions (Cross-Chapter Box

                                                                                                       URBAN in Chapter 6)

Do Not Cite, Quote or Distribute  18-100                       Total pages: 197
FINAL DRAFT                       Chapter 18                  IPCC WGII Sixth Assessment Report

 Health,        SDG3,   � a lack of capacity for adaptation   � urban ecological infrastructure    � cooling networks (Cross-          Heat Island (UHI) effect (Cross-
 wellbeing and  SDG5,     has resulted in only moderate or       can be a source of nature-based                                       Chapter Box URBAN in Chapter
 the changing   SDG8,     low levels of adaptation               solutions that can improve both     Chapter Box URBAN in              6)
 structure of   SDG10,    implementation across different        adaptation and mitigation in                                        � (T) efforts aimed at increasing
 communities    SDG13     countries (7.4.2)                      urban areas (Cross-Chapter Box      Chapter 6)                        adaptation may undermine
                                                                 URBAN in Chapter 6)               � early warning systems (Table      mitigation objectives by increasing
                        � transitioning to renewable energy                                                                            investment in hard infrastructure
                          sources presents opportunities for  � high density environments            6.4)                              that increases emissions (Cross-
                          realizing health co-benefits           coupled with other design         � resource demand and supply        Chapter Box URBAN in Chapter
                          (7.4.4)                                measures can provide mitigation                                       6)
                                                                 and adaptation benefits (Cross-     side management strategies      � (T) lack of open and green spaces
                        � shifting to healthier plant-rich       Chapter Box URBAN in                (Table 6.4)                       may induce long-distance leisure
                          diets can reduce GHG emissions         Chapter 6)                        � enhanced monitoring of air        trips thereby increasing emissions
                          and reduce land-use (Cross-                                                                                  and (Cross-Chapter Box URBAN
                          Chapter Box HEALTH in               � COVID-19 recovery investments        quality in rapidly developing     in Chapter 6)
                          Chapter 7)                             offer an opportunity to
                                                                 contribute to climate resilient     cities (Table 6.4)              � (T) energy strategies for energy
                        � future flows of migration within       development through a green,      � investment in air pollution       efficiency and GHG emissions
                          and between countries are likely       resilient, healthy and inclusive                                      reductions can generate health co-
                          to respond strongly to particular      recovery (Cross-Chapter Box         controls (Table 6.4)              benefits through improved air
                          combinations of climatic hazards       COVID in Chapter 7)               � core and shell preservation,      quality but may slow poverty
                          and may present challenges for                                                                               reduction efforts (7.4.2, 7.4.5)
                          future adaptation policies and      � investing in basic infrastructure    elevation and relocation for
                          programs                               for all can transform                                               � (S+) investing in adaptation for
                                                                 development opportunities,          heritage buildings (6.3.2)        health and community wellbeing
                        � climate change disruptions to          increase adaptive capacity and    � improved building and urban       has the potential to generate
                          natural environments can be            reduce climate risk (Cross-                                           considerable co-benefits in terms
                          expected to disrupt livelihood         Chapter Box HEALTH in               design including use of           of reducing impacts of non-climate
                          practices, stimulate higher rates      Chapter 7)                                                            health challenges
                                                                                                     passive cooling systems
                                                              � Integrated agroecological            (Table 7.2)                     � (S+) investments in mitigating
                                                                 systems offer opportunities to    � better access to public health    greenhouse gas emissions will not
                                                                 increase dietary diversity while                                      only reduce risks associated with
                                                                 building local resilience to        systems for the most              dangerous climate change, but will
                                                                 climate-related food insecurity                                       increase population health and
                                                                 (7.4.2)                             vulnerable (Table 7.2)            wellbeing through a number of
                                                                                                   � deployment of renewable           pathways. (7.4)
                                                              � Incorporating climate change
                                                                 and health considerations into      energy sources (Table 7.2)
                                                                 disaster reduction and            � improved water, sanitation
                                                                 management strategies could
                                                                                                     and hygiene conditions

                                                                                                     (Table 7.2)
                                                                                                   � early-warning system of

                                                                                                     vector-borne diseases,
                                                                                                     insecticide treated bed nets,

                                                                                                     and indoor spraying of

                                                                                                     insecticide (Table 7.2)
                                                                                                   � targeted efforts to develop

                                                                                                     vaccines for infectious

                                                                                                     diseases exacerbated by

                                                                                                     climate change (Table 7.2)
                                                                                                   � improved personal drinking

                                                                                                     and eating habits (Table 7.2)

Do Not Cite, Quote or Distribute  18-101                      Total pages: 197
FINAL DRAFT                       Chapter 18                 IPCC WGII Sixth Assessment Report

 Poverty,     SDG1,      of outmigration to urban centers,     potentially improve funding        � improved food storage, food      � (S+) agriculture technologies
 livelihoods  SDG2,      and in some instances necessitate     opportunities (7.4.2)                                                   facilitate mitigation to climate
 and          SDG3,      planned or organized relocations    � adaptive urban design that           processing, and food               change and adaptation such as
 sustainable  SDG5,      of exposed settlements (Cross-        provides access to healthy                                              saving water while maintaining
 development  SDG10,     Chapter Box MIGRATE in                natural spaces can promote           preservation (Table 7.2)           grain yield (8.6.1)
              SDG14      Chapter 7)                            social cohesion and mitigate       � emergency shelters for people
                                                               mental health challenges (7.4.2)                                      � (S+) sustainable pastoralism
                      � use of political frameworks for                                             to escape heat (Table 7.2)         increases carbon sequestration but
                        decision-making that are             � polycentric governance, adaptive   � improved funding and access        can also contribute to adaptation
                        unfavorable towards adaptation         governance, multi-level                                                 by changing grazing management,
                        and system transitions (Table 8.4)     governance, collaborative            to mental health care (Table       livestock breeds, pest management,
                                                               governance, or network                                                  and production structures (8.6.1)
                      � attitudes toward risk and other        governance are increasingly used     7.2)
                        cultural values limit responses        to understand transitions towards  � improved education for girls     � (S+) REDD+ may provide
                        (Table 8.4)                            climate-compatible development                                          adaptation benefits by enhancing
                                                               (8.6.2)                              and women (Table 7.2)              households' economic resilience
                      � psychological distress causes                                             � improved maternal and child        through positive livelihood impacts
                        insecurity and behaviors that        � well-coordinated and integrated                                         (8.6.1)
                        increase vulnerability (Table 8.4)     nexus approaches to adaptation       health services (Table 7.2)
                                                               offer opportunities to build       � expanded private sector          � (S+) solar energy contributes to
                      � limited financial resources to         resilient systems while                                                 reducing GHG emissions and
                        support adaptation projects (8.2.2,    harmonizing interventions,           activity and public-private        improving air quality (8.6.1)
                        Table 8.4)                             mitigating trade-offs and
                                                               improving sustainability (8.6.2)     partnerships (8.6.1)             � (S+) hydropower contributes to
                      � small-holder farmers have poor                                            � credit and insurance (8.6.1)       mitigation and adaptation through
                        access to markets and land tenure    � income from new livelihood         � use of climate-smart               water resource availability for
                        (8.6.1)                                activities can support recovery                                         irrigation and drinking water
                                                               following disasters linked to        agricultural practices and         (8.6.1)
                      � unsuitable infrastructure may          climate variability and change
                        increase exposure (Table 8.4)          (8.4.5)                              technologies (8.6.1)             � (S+) green roofed buildings
                                                                                                  � crop insurance (8.6.1)             contribute to cooler temperatures,
                      � lack of access to technologies that  � improving industrial processes     � conservation agriculture           thereby reducing energy use for
                        can support adaptation (Table          can contribute to the optimized                                         air-conditioning (8.6.1)
                        8.4)                                   use of energy, reuse of waste,       (8.6.1)
                                                               reducing GHG emissions, use of     � changing farmers' perception
                      � gender-based inequalities              biomass and more efficient
                        constrain women's access to            equipment (Table 8.3)                and enhancing farmers'
                        resources for adaptation (Table
                        8.7)                                 � industrialization and                adaptive capacity (8.6.1)
                                                               technological innovation in rural  � REDD+ (8.6.1)
                      � poverty constrains livelihood          areas may assist vulnerable        � improving industrial
                        diversification, resilience or
                        adaptive capacity (Table 8.7)                                               processes (Table 8.3)
                                                                                                  � renewable energy and energy

                                                                                                    efficiency (Table 8.3)
                                                                                                  � smart electricity grids (8.6.1)
                                                                                                  � green buildings (8.6.1)
                                                                                                  � efficient fuels (8.6.1)
                                                                                                  � pollution control investments

                                                                                                    (8.6.1)
                                                                                                  � public transit and non-

                                                                                                    motorized transport with

                                                                                                    increased use of biofuels

                                                                                                    (8.6.1)

Do Not Cite, Quote or Distribute  18-102                     Total pages: 197
      FINAL DRAFT                 Chapter 18             IPCC WGII Sixth Assessment Report

1                  � indigenous peoples and other          communities through provision      � integrated natural resource    � (T) mitigation measures such as
                     populations with strong               of resources, enhanced forecast                                       bioenergy may result in trade-offs
                     attachments to place face barriers    information, or reuse of biowaste    management (Table 8.2)           with efforts to achieve sustainable
                     to adaptation (Table 8.7)             (Table 8.3)                        � disaster risk management         development, eradicate poverty
                                                         � responses to climate change can                                       and reduce inequalities (8.6.1)
                   � local institutions face ongoing       create significant development       (Table 8.2)
                     challenges in gaining support         opportunities including job        � relocation of vulnerable       � (T) migration to urban centers can
                     from higher governance levels,        creation and livelihood                                               be a form of adaptation, but can
                     particularly in developing            diversification (8.4.3)              communities (Table 8.2)          increase the vulnerability of
                     countries. (8.5.2)                                                       � Education and                    communities of origin or at
                                                                                                                                 destinations (8.2.2)
                                                                                                communication (Table 8.2)
                                                                                              � land use planning (Table 8.3)

Do Not Cite, Quote or Distribute  18-103                 Total pages: 197
    FINAL DRAFT                                      Chapter 18  IPCC WGII Sixth Assessment Report

1 18.6 Conclusions and Research Needs

2

3 18.6.1 Knowledge Gaps

4

5 Research to improve the understanding of CRD currently exists in a nascent state, because, as noted in the
6 AR5, "integrating climate change mitigation, climate change adaptation, and sustainable development is a
7 relatively new challenge" (Denton et al., 2014). While a large volume of literature has emerged since the

8 AR5 that spans the nexus of sustainable development, CRD, and climate action, the identified research gaps

9 in AR5 (Denton et al., 2014) continue to be priorities for informing CRD. These include enhancing

10 understanding of mainstreaming of climate change into institutional decision-making, managing risk under

11 conditions of uncertainty, catalyzing system transitions and transformation, and processes for enhancing
12 participation, equity, and accountability in sustainable development (very high confidence).

13

14 The more recent literature adds significant context to the concept of CRD, but also introduces broader

15 perspectives regarding its significance in the arena of climate action. Hence, concepts that are both

16 complementary to, and competitive with, CRD, such as climate safe', `climate compatible' and `climate

17 smart' development (Huxham et al., 2015; Kim et al., 2017b; Ficklin et al., 2018; Mcleod et al., 2019)

18 (18.1.1). These different framings of the intersection between sustainable development and climate action

19 are used in different communities of research and practice, which complicates efforts to provide clear

20 guidance to decision-makers regarding the goals of CRD and how best to achieve it. This is attributable in

21 part to persistent conceptual confusion and disciplinary divides over more fundamental concepts such as

22 resilience and sustainability (Rogers et al., 2020; Zaman, 2021), not to mention contested perspectives
23 regarding development (Lo et al., 2020; Song et al., 2020a; Morton, 2021) (medium agreement; medium
24 evidence).

25

26 Reconciling different perspectives on CRD is not simply a matter of academic debate. Climate action,

27 resilience, and sustainable development are all active areas of policy and practice with significant economic,

28 social, environmental, and political implications (18.1.3). Hence, enhancing the role of CRD as a practical

29 framework for development and a guide for action may necessitate improving the science-policy discourse

30 regarding CRD (Winterfeldt, 2013; Jones et al., 2014; Ryan and Bustos, 2019). This includes consideration

31 for risk and science communication; decision analysis and decision support systems; and mechanisms for
32 knowledge co-production between scientists and public policy actors (very high confidence).

33

34 In addition, the AR6 WGII report highlights a number of elements of CRD that are associated with

35 significant knowledge gaps and uncertainties. As a result, enhancing the value of CRD as a unifying concept

36 in development would benefit from further conceptualization and socialization of the concept as well as

37 efforts to address the following knowledge gaps:

38  � The challenges posed by different levels of global warming to achieving CRD and the magnitude

39  and nature of the adaptation gap (and associated finance needs) that must be addressed to enable

40  climate resilience.

41  � The efficacy of different adaptation, mitigation, and sustainable development interventions in

42  reducing climate risk and/or enhancing opportunities for CRD in the short, medium and long term.

43  � How different CRD pathways can be designed such that they illustrate opportunities for the practical

44  pursuit of CRD in a manner consistent with principles of inclusion, equity, and justice.

45  � How deliberative, participatory learning can be integrated into approaches to CRD in order to

46  enhance the representation of diverse actors, forms of knowledge, governance regimes, economic

47  systems, and models for decision-making in CRD.

48  � The synergies and trade-offs associated with the implementation of different policy packages and the

49  design principles and development contexts that enhance the ability to successfully manage potential

50  trade-offs.

51  � The limits of incremental system transitions to achieving CRD on a timeline that reflects the urgency

52  associated with the Paris Agreement and the Sustainable Development Goals.

53  � The capacity of governments, social institutions, and individuals to drive large-scale social

54  transformations that open up the solutions space for CRD.

    Do Not Cite, Quote or Distribute                 18-104                     Total pages: 197
   FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1  � Best practices for avoiding maladaptation and ensuring that adaptation interventions are designed so

2  they do not exacerbate vulnerability to climate change to support CRD.

 3 18.6.2 Conclusions

 4

 5 The concept of CRD presents an ambitious agenda for actors at multiple scales � global to local, particularly
 6 in the manner in which it reframes climate action to integrate a broader set of objectives than simply
 7 reducing greenhouse gas emissions or adapting to the impacts of climate change. Specifically, recent
 8 literature extends policy goals for climate action beyond avoiding dangerous interference with the climate
 9 system to adopt normative goals of meeting basic human needs, eliminating poverty and enabling sustainable
10 development in ways that are just and equitable. This creates a policy landscape for climate action that is not
11 only richer, but also more complex in that it situates responses to climate change squarely within the
12 development arena. Current policy goals associated with the Paris Agreement, Sendai Framework, and the
13 SDGs imply aggressive timetables. Yet, as noted in the AR5 and supported by more recent literature
14 (Section 18.2.1), the world is neither on track to achieve all of the SDGs nor fulfil the Paris Agreement's
15 objective of limiting warming to well-below 2癈 (Denton et al., 2014; IPCC, 2018a). This places aspirations
16 for CRD in a precarious position. Transitions will be necessary across multiple systems (Section 18.1.3).
17 While some may be already underway, the pace of those transitions must accelerate, and societal
18 transformations may be necessary, to enable CRD (18.3, 18.4, Box 18.1)

19

20 Given the pace of climate change and the inherent challenge of sustainable development, particularly in the
21 face of inevitable disruptions and setbacks such as the COVID-19 pandemic (Cross-Chapter Box COVID in
22 Chapter 7), the feasibility of achieving CRD is an open question. Rapid changes will be required to shift
23 public and private investments, strengthen institutions and orient them toward more sustainable policies and
24 practices, expand the inclusiveness of governance and the equity of decision-making, and shift societal and
25 consumer preferences to more climate-resilient lifestyles. Nevertheless, the collective body of recent
26 literature on CRD, system transitions, and societal transformation, combined with the assessments within
27 recent IPCC Special Reports (IPCC, 2018a; IPCC, 2019b; IPCC, 2019d) indicate that there are a broad range
28 of opportunities for designing and implementing adaptation and mitigation options that enable the climate
29 goals in the Paris Agreement to be achieved while enhancing resilience and meeting sustainable
30 development objectives. However, options should be considered alongside the mechanisms by which
31 societies can engage in order to create the conditions that can support the implementation of those options
32 (Section 18.4). This includes formal policy mechanisms pursued by governments, the catalyzation of
33 innovation by private firms and entrepreneurship, as well as informal, grassroots interventions by civil
34 society. While there is no "one-size-fits-all" solution for CRD that will work for all actors at all scales,
35 exploring different pathways by which actors can achieve their development and climate goals can make
36 valuable contributions to developing effective strategies for CRD.

37

38 A fundamental challenge for achieving CRD globally is reconciling different perspectives on CRD. As noted
39 in the AR5, "as policy makers explore what pathways to pursue, they will increasingly face questions about
40 managing discourses about what societal objectives to pursue" (Denton et al., 2014: 1124). Since the AR5,
41 such discourses have become prominent in policy debates over climate action and sustainable development
42 due to different nations, communities, and subpopulations having different understandings of what
43 constitutes CRD. Aggressive efforts to rapidly reduce greenhouse gas emissions or enhance resilience to
44 climate change, for example, could have negative externalities for the development objectives of some
45 actors. This potential for trade-offs complicates efforts to build consensus regarding what constitutes
46 appropriate climate and development policies and practices and by whom. The CRD pathways preferred by
47 one actor are likely to be contested by others. This means operationalizing concepts such as CRD in practice
48 is likely to necessitate ongoing negotiation.

49

50 Ultimately, one of the critical developments within the literature is the emergence of procedural and
51 distributive justice as key criteria for evaluating climate action and CRD more specifically. This trend not
52 only recognizes the need to prevent vulnerable human and ecological systems from experiencing
53 disproportionate harm from the changing climate, but also the need to prevent those same systems from
54 being harmed by mitigation, adaptation, and sustainable development policies and practices. Failure to
55 adequately engage with equity and justice when designing sustainability transitions could lead to
56 maladaptation, aggravated poverty, reinforcement of existing inequalities, and entrenched gender bias and

   Do Not Cite, Quote or Distribute  18-105                                Total pages: 197
    FINAL DRAFT                       Chapter 18                   IPCC WGII Sixth Assessment Report

1 exclusion of Indigenous and marginalized communities (Jenkins et al., 2018; Fisher et al., 2019; Schipper et

2 al., 2020b). These consequences could ultimately slow, rather than accelerate, CRD. Hence, developing

3 programs and practices for prioritizing equity in effective transition risk management is an important

4 dimension of enabling CRD.

5

6 As indicated by the literature assessed within this chapter, keeping windows of opportunity open for CRD

7 will necessitate urgent action, even under diverse assumptions regarding how future mitigation and

8 adaptation interventions evolve. If nations are to collectively limit warming to well-below 2癈, for example,

9 unprecedented emissions reductions will be necessary over the next decade (IPCC, 2018a). These reductions

10 would necessitate rapid progression of system transitions (18.3). If, despite the Paris Agreement, future

11 emissions trajectories take the world beyond 2癈, a greater demand will be placed on adaptation as a means

12 of enhancing the resilience of development. Given the long-lived nature of human systems, and the built

13 environment in particular, significant adaptation investments would be needed over the near-term to meet

14 this demand. Yet, it is important to note that even in the absence of consideration for climate change,

15 substantial development needs exist for communities around the world at present. Hence, a robust strategy

16 for the pursuit of CRDPs is a near-term focus on portfolios of policies and practices that promote of human

17 and ecological well-being.

18

19 [START FAQ18.1 HERE]

20

21 FAQ18.1: What is a climate resilient development pathway?

22

23 Climate resilient development pathways (CRDPs) are continuous processes that strengthen sustainable

24 development, efforts to eradicate poverty and reduce inequalities while promoting fair and cross-scalar

25 capacities for adaptation to global warming and reduction of greenhouse gases in the atmosphere.

26

27 A pathway is defined in IPCC reports as a temporal evolution of natural and/or human systems towards a

28 future state. These can range from sets of scenarios, narratives of potential futures to solution-oriented

29 decision-making processes to achieve desirable societal goals.

30

31 When used in the context of climate resilient development (CRD), pathways refer to continuous processes

32 that strengthen sustainable development, efforts to eradicate poverty, and reduce inequalities while

33 promoting fair and cross-scalar adaptation and mitigation. As they imply deep societal changes and/or

34 transformation, CRDPs raise questions of ethics, equity, and feasibility of options to drastically reduce

35 emission of greenhouse gasses (mitigation) that limit global warming (e.g., to well below 2癈) and achieve

36 desirable and livable futures and wellbeing for all.

37

38 There in no one true, correct pathway to pursue but multiple ways, modalities, depending on numerous

39 factors, such as political, cultural and economic contexts. Pathways are not one single decision or action, nor

40 is there an absolute, universal, fixed, final goal to be pursued, yet there are undesirable and non-CRDPs.

41 Hence, a CRDP is a continuum of coherent, consistent decisions, actions and interventions within each

42 country, and as a global community. While dependent on past development and its socio-ethical, political,

43 economic, ecological and knowledge-technology outcomes at any point in time, transformation, ecological

44 tipping points and shocks can create sudden shifts and unexpected non-linear development pathways.

45 Actions taken today also foreclose some future potential pathways. The differentiated impacts of hurricanes

46 and COVID-19 illustrate how the character of societal development such as equity and inclusion have

47 enabled some societies to be more resilient than others.

48

49 [END FAQ18.1 HERE]

50

51

52 [START FAQ18.2 HERE]

53

54 FAQ18.2: What is climate resilient development and how can climate change adaptation (measures)

55  contribute to achieving this?

56

    Do Not Cite, Quote or Distribute                     18-106    Total pages: 197
FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

 1 The key purpose of CRD is to pursue sustainable development, engaging climate actions in ways that
 2 support human and planetary health and well-being, equity and justice. Climate resilient development
 3 combines adaptation and mitigation with underlying development choices and everyday actions, carried out
 4 by multiple actors within political, economic, ecological, socio-ethical and knowledge-technology arenas.
 5 The character of processes within these development arenas are intrinsic to how social choices are made,
 6 directing actions in a CRD or non-CRD direction. For example, inclusion, agency and social justice are
 7 qualities within the political arena that underpin actions that enable CRD.

 8

 9 CRD addresses the relationship between greenhouse gas emissions, levels of warming and related climate
10 risks. However, CRD involves more than just achieving temperature targets. It considers the possible
11 transitions that enable those targets to be achieved as well as the evaluation of different adaptation strategies
12 and how the implementation of these strategies interact with broader sustainable development efforts and
13 objectives. This interdependence between patterns of development, climate risk, and the demand for
14 mitigation and adaptation action is fundamental to the concept of CRD. Therefore, climate change and
15 sustainable development cannot be assessed or planned in isolation of one another.

16

17 Hence, CRD is defined as the development that deliberately adopts mitigation and adaptation measures to
18 secure a safe climate on earth, meet basic needs for each human being, eliminate poverty and enable
19 equitable, just and sustainable development. It halts practices causing dangerous levels of global
20 warming. CRD may involve deep societal transformation to ensure well-being for all. CRD is now emerging
21 as one of the guiding principles for climate policy, both at the international level, reflected in the Paris
22 Agreement (UNFCCC, 2015) and within specific countries.

23
24

25

26 Figure FAQ18.2.1: Multiple intertwined climate resilient development pathways. Climate change adaptation is one
27 of several climatic and non-climatic measures carried out through decision-making by multiple actors that may drive a
28 pathway in a CRD or non-CRD direction. Adaptation, mitigation and sustainable development actions can push a
29 society in a CRD direction, but only if these measures are just and equitable. There are multiple simultaneous pathways
30 in the past, present and future. Societies (illustrated as boats) move on different pathways, towards CRD and non-CRD,
31 with some pathways more dominant than others. The direction of pathways is emergent, taking place through
32 contestations and social choices, through social transformation as well as through surprises and shocks (illustrated as
33 rocks). Path dependency means it is possible but often turbulent to shift from a non-CRD to a CRD pathway. Such a
34 shift becomes more difficult in as risks/shocks increase (more rocks) and non-CRD processes and outcomes progress,
35 limiting future options. Low CRD processes and outcomes at the bottom are characterized by inequity, exclusion,
36 polarization, environmental and social exploitation, entrenchment of business as usual, with increasing risks/shocks.
37 High CRD processes and outcomes (at the top of the figure) are characterized by equity, solidarity, justice, human well-
38 being, planetary health, stewardship/care and system transitions.

39

Do Not Cite, Quote or Distribute  18-107      Total pages: 197
    FINAL DRAFT                       Chapter 18  IPCC WGII Sixth Assessment Report

1 Climate change adaptation is one of several climatic and non-climatic measures carried out through decision-

2 making by multiple actors that may drive a pathway in a CRD or non-CRD direction. Adaptation, mitigation

3 and sustainable development actions can push a society in a CRD direction, but only if these measures are

4 just and equitable. There are multiple simultaneous pathways in the past, present and future. Societies

5 (illustrated as boats) move on different pathways, towards CRD and non-CRD, with some pathways more

6 dominant than others. The direction of pathways is emergent, taking place through contestations and social

7 choices, through social transformation as well as through surprises and shocks (illustrated as rocks). Path

8 dependency means it is possible but often turbulent to shift from a non-CRD to a CRD pathway. Such a shift

9 becomes more difficult in as risks/shocks increase (more rocks) and non-CRD processes and outcomes

10 progress, limiting future options. Low CRD processes and outcomes at the bottom are characterized by

11 inequity, exclusion, polarization, environmental and social exploitation, entrenchment of business as usual,

12 with increasing risks/shocks. High CRD processes and outcomes (at the top of the figure) are characterized

13 by equity, solidarity, justice, human well-being, planetary health, stewardship/care and system transitions.

14

15 [END FAQ18.2 HERE]

16

17

18 [START FAQ18.3 HERE]

19

20 FAQ18.3: How can different actors across society and levels of government be empowered to pursue

21  climate resilient development?

22

23 CRD entails trade-offs between different policy objectives. Governments, political and economic elites may

24 play a key role in defining the direction of development at a national and sub-national scale; but in practice,

25 these pathways can be influenced and even resisted by local people, NGOs and civil society.

26

27 Contestation and debate are inherent in its construct and implementation. An active civil society and

28 citizenship create the enabling conditions for deliberation, protest, dissent and pressure which are

29 fundamental for an inclusive participatory process. These enable a multiplicity of actors to engage across

30 multiple arenas, from decision-making and everyday actions Hence, decisions and actions may be influenced

31 by uneven interactions between actors, including socio-political relations of domination, marginalization,

32 contestation, compliance and resistance with diverse and often unpredictable outcomes.

33

34 In this way, recent social movements and climate protests show new modalities of action related to political

35 responsibility for inaction based on contestation. The new climate movement led mostly by youngsters,

36 markedly seek science-based policy and more importantly, demand to break with a reformist stance and

37 social inertia through radical climate action. This is mostly done through collective disruptive action, and

38 non-violent resistance to promote awareness, a regenerative culture and ethics of care. These movements

39 have resulted in notable political successes, such as declarations of climate emergency at the national and

40 local level, as well as in universities. Also, their methods have proven effective to end fossil fuel

41 sponsorship.

42

43 The success and importance of recent climate movements also provide elements to rethink the role of science

44 in society. In one hand, the new climate movements demanding political action were prompted by the

45 findings of scientific reports, mainly the IPCC (2018a) and IPBES (2019) reports. On the other hand, these

46 movements have increased public awareness, and also stimulated public engagement with climate change at

47 unprecedented levels.

48

49 [END FAQ18.3 HERE]

50

51

52 [START FAQ18.4 HERE]

53

54 FAQ18.4: What role do transitions and transformations in energy, urban and infrastructure,

55  industrial, land and ocean ecosystems, and in society, play in climate resilient development?

56

    Do Not Cite, Quote or Distribute  18-108                                                    Total pages: 197
    FINAL DRAFT                                       Chapter 18  IPCC WGII Sixth Assessment Report

1 The IPCC 1.5 report identified transitions and transformations in key systems, such as energy, land, and

2 ocean ecosystems, and urban and infrastructure, that are needed for a climate resilient development. A

3 system transitions focus helps visualize the interdependence between each system as well as how sustainable

4 development, mitigation, and adaptation interact. A societal transformation, in terms of values and

5 worldviews that shape aspirations, lifestyles and consumption patterns, is a constraining/enabling condition

6 for such transformations. This report however identifies societal transformation as one of the five major

7 transformations currently underway. It delves into the implications of this on how we assess options, value

8 different outcomes from the perspectives of ethics, equity, justice and inclusion.

9

10 [END FAQ18.4 HERE]

11

12

13 [START FAQ18.5 HERE]

14

15 FAQ18.5: What are success criteria in climate resilient development and how can actors satisfy those

16  criteria?

17

18 Climate resilient development is not a predefined goal to be achieved at a certain point or stage in the future.

19 It is a constant process of evaluating, valuing, acting and adjusting various options for mitigation, adaptation

20 and sustainable development, shaped by societal values as well as contestations of these. Any achievement

21 or success is always a work in progress, with continuous, directed, intentional actions. These actions will

22 vary according to the priorities and needs of each population or system; therefore, specific indicators will

23 vary according to each specific context, ensuring we prioritize people, planet, prosperity, peace, and

24 partnership, per the broad goals of the Agenda 2030 on sustainable development.

25

26 If Climate Resilient Development is defined as the development that deliberately adopts mitigation and

27 adaptation measures to secure a safe climate, meet basic needs, eliminate poverty and enable equitable, just

28 and sustainable development, then, the 17 United Nations' Sustainable Development Goals (SDGs) provide

29 a good (although limited) measure of progress. They aim at ending poverty and hunger globally and protect

30 life on land and under water until the year 2030. Although there are proven synergies between the SDGs and

31 mitigation, there remains to explore clear synergies between the SDGs and adaptation in terms of how

32 adaptation relates to the fulfilment of the SDGs.

33

34

    Do Not Cite, Quote or Distribute                  18-109                          Total pages: 197